You are on page 1of 12

J Mater Sci

The Physics of Metal Plasticity: In honor of Professor Hussein Zbib

Evolution of residual stress and interface coherency


and their impact on deformation mechanisms in Al/
Ti multilayers
Wenbo Wang1, Sina Izadi2, Hesham Mraied2, Chuang Deng3,*, and Wenjun Cai1,*

1
Department of Materials and Science Engineering, Virginia Polytechnic Institute and State University, Blacksburg, VA 24061, USA
2
Department of Mechanical Engineering, University of South Florida, Tampa, FL 33620, USA
3
Department of Mechanical Engineering, University of Manitoba, 75A Chancellors Circle, Winnipeg, MB R3T 5V6, Canada

Received: 15 July 2023 ABSTRACT


Accepted: 29 August 2023 Hardness of nanostructured metallic multilayers (NMMs) are often understood
from their dependence on individual layer thickness (h), yet little is known about
© The Author(s), under the impacts exerted by other microstructural factors. In this work, the effects of
exclusive licence to Springer residual stress and interface coherency on the deformation mechanisms of Al/Ti
Science+Business Media, LLC, NMMs with h = 2.5–52 nm were studied via experiments and molecular dynamics
part of Springer Nature, 2023 simulations. The residual stress was found to be tensile in Ti layers and compres-
sive in Al layers in general, both of which tended to increase with decreasing h.
Tensile stress of more than 2 GPa were measured in the Ti layers at h < 10 nm.
Such high stress was related to the more coherent interfaces at small h, as con-
firmed by transmission electron microscopy analysis. Finally, molecular dynamics
simulations showed that at h = 2.5 nm, the coherent interfaces were more effective
barriers to dislocation initiation and transfer than the incoherent ones.

Introduction heterogeneity on the surface [4]. The high interfacial


area between the layers provides enhanced mechani-
Nanostructured metallic multilayers (NMMs) con- cal strength, improved thermal stability, and modi-
sisting of alternating layers of different metals or fied electrical and optical properties compared to
alloys have attracted significant interests due to their their bulk counterparts. NMMs find applications in
potential to achieve improved mechanical, optical and various fields, including microelectronics, catalysis,
electrochemical properties compared to monolithic sensors, and coatings, due to their tailored properties
coatings of equivalent thickness [1–3]. Such layered and potential for improved performance in specific
structure can strengthen the material by confining the applications.
dislocation movements via Hall–Petch and related NMMs exhibit a well-known length-scale-depend-
mechanisms without causing unfavorable chemical ent strengthening and deformation mechanism

Handling Editor: Tariq Khraishi.

Address correspondence to E-mail: Chuang.Deng@umanitoba.ca; caiw@vt.edu

https://doi.org/10.1007/s10853-023-08908-3

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Vol.:(0123456789)
J Mater Sci

behavior [5–7], which has been studied using dislo- stress and coating strength [14–16]. For example,
cation-mediated plasticity models such as disloca- positive (or tensile) residual stress in monolithic Ti
tion pile-up model, confined layer slip (CLS) model reduces the overall strength of the coating [17] and
etc. [7, 8]. The deformation behavior depends on the negative (compressive) residual stress increases the
thickness of the layer (h), resulting in three distinct strength of Al coating [18].
modes [7, 9, 10]. When the thickness falls within the Measuring the residual stress is often challenging
submicron range, dislocations accumulate at the inter- experimentally. For monolithic thin films, the macro-
faces and the strength increases as the layer thickness scopic stress or strain can be measured relative easily
decreases following the Hall–Petch relationship. In using the curvature method by a stylus profilometry
the range of a few to a few tens of nanometers, the or laser and white-light interferometry. The curvature
interfaces are too close for dislocation pile-up to developed of the film is then used to calculate the
occur, leading to the breakdown of Hall–Petch hard- residual stress in the whole film. The residual stress
ening. Instead, the deformation mechanism involves or strain in NMMs needs to be measured within indi-
the movement of individual dislocations bending vidual layers, which is highly sensitive to interface
between the interfaces, known as confined layer slip coherency and individual layer thickness. In addition,
(CLS) mechanism. In this regime, the yield strength a complete 3D representation of the stress field, which
increases with decreasing layer thickness, following a requires 6 independent stress components, is often not
relationship of 𝜎 ∝ ln (h)∕h [7]. At extreme small layer reported. In thin film NMMs, residual stress is most
thickness (h < 10 nm), the multilayers are also known often reported as a scalar, corresponding to the normal
as superlattices. In the regime, the strength become stress component in the film growth direction, while
less dependent on layer thickness but highly sensitive normal stress on the film plane direction or any shear
to the structure and properties of the interfaces such component are hard to measure.
as coherency strains [11]. X-ray diffraction using the ­sin2ψ method is a widely
Unlike layer thickness or interface coherency, the used technique to determine the residual stress in thin
effect of residual stress on the mechanical properties films [19], where Ψ is the angle between incident-dif-
of NMMs is less understood [12]. NMM are often fracted bisector and the normal of sample surface. For
manufactured via thin film deposition methods such example, Ljungcrantz et al. [20] showed that during dc
as physical vapor deposition or electrodeposition. magnetron sputtering, the residual stress developed
During the deposition process, atoms or molecules in Ti thin films deposited on Si substrate changes
from the vapor phase condense and form layers. from ~ 0.4 GPa tensile to ~ 0.4 GPa compressive as
The growth of each layer can induce stress due to the argon pressure decreased below 2.5 mTorr. Staf-
lattice mismatch, defects, or strain accumulation. If ford et al. [21] found that in electrodeposited Al thin
heating was applied to the substrate or environment films on Au substrate, the residual stress was initially
during the deposition process, due to differences in compressive at small film thickness (> ~50–100 nm),
coefficients of thermal expansion (CTE) between and quickly becomes tensile (~ 25–75 MPa) at larger
the alternative layered materials and the substrate, thicknesses due to lattice mismatch between Al and
additional residual stress would develop when the Au as well as AlAu alloying. In terms of multilayers,
film cools down to room temperature. Earlier work Gu et al. [22] measured the residual strain field of
also showed that the magnitude of residual stress CrN/AlN multilayers with bilayer period of 6.0, 5.5,
is highly sensitive to the overall film thickness. For and 2.0 nm. They concluded that a tensile strain of
example, the residual stress changes through Vol- ~ 0.007–0.015 develops in the CrN layers while a com-
mer–Weber growth mechanism from compressive to pressive strain of ~ − 0.013 to 0.017 develops in the
tensile and then compressive again while the layer AlN layers. Recently, Todt et al. [23] performed in-situ
thickness is increased. Due to the ultra-fine layer nanoindentation of CrNi-AlN superlattice thin films
thickness, high levels of residual stress can develop using synchrotron X-ray nanoprobe. They found that
in superlattice coatings (i.e., NMMs with h< 10 nm). a very high compressive stress of ~ 13 GPa developed
For example, residual stress of Au layers (h = 9 Å) under the indent, followed by a recovery by tensile
in Au/Ni NMMs approaches ~ 3.9 GPa, significantly stress of ~ 1.4 GPa.
higher than the yield stress of bulk Au [13]. Past stud- Despite prior studies, the reports on the measure-
ies indicate a direct relationship between the residual ment of residual stress and interface coherency of

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

NMMs are still lacking. Hence, the purpose of this were chosen due to obviously distinguish two phases
work is to (1) measure the evolution of residual stress instead of superposition at ~ 38.5° and measured at Ψ
in Al/Ti NMMs with different h, (2) determine the tilt angles of 0°, ± 10°, ± 15°, ± 30° and ± 35°. The sur-
effects of layer thickness on the interface coherency, face morphology of all samples was characterized by
and (3) quantify the effects of residual stress/strain Hitachi Su-70 scanning electron microscopes (SEM).
and interface coherency on the deformation behavior FEI Tecnai G2 transmission electron microscopy (TEM)
and hardness of NMMs. Prior studies on NMMs often was applied prior and post nanoindentation for char-
focus on FCC/FCC or FCC/BCC types, Al/Ti NMMs is acterization of samples under 200 kV acceleration volt-
chosen here to represent an FCC/HCP system, which is age. TEM samples were prepared by FEI quanta 200
less studied. In this work, experimental measurements 3D focused ion beam (FIB) using the standard lift-out
and molecular dynamics simulations are integrated to method. Before milling, a protective carbon layer with
shed light on the design of future high strength NMMs a thickness of 30–100 nm was first deposited on sam-
via engineering of internal strains and interfaces. ples, followed by Pt deposition up to 200 nm using
electron beam and 2–3 μm using ion beam. To meas-
ure their mechanical properties and induce deforma-
Experimental and computational methods tion behaviors of Al/Ti NMMs, nanoindentation tests
using Hysitron Ti900 Triboindenter with a new stand-
Equal-spaced Al/Ti NMMs with different h from 52 to ard Berkovich tip were carried out and a more detailed
2.5 nm, as well as monolithic Al and Ti thin films, were description of parameters can be found in [24].
deposited on Si (100) substrate using a SFI sputtering Molecular dynamics (MD) was also used to simulate
tool with an Eratron 8210 DC power supply, as listed the nanoindentation of Al/Ti NNMs by using Large-
in Table 1, following the procedure described in [24]. scale Atomic/Molecular Massively Parallel Simulator
In all NMMs, the Al layer is always the topmost layer. (LAMMPS) package [26] with an embedded-atom
One of the reasons for selecting Al as the topmost layer method potential for the Al–Ti system [27]. As shown
is due to its good corrosion resistance, as described in in Fig. 1, a three-dimensional slab model was used to
our previous work [25]. PANalytical X’Pert PRO dif- model the Al/Ti NNMs with fixing sizes along X, Y
fractometer (Cu Kα, 1.54 Å) was applied to conduct and Z, i.e., ­LX ~ 100.5 nm, ­LY ~ 1.5 nm, ­LZ ~ 63 nm while
grazing incident XRD (GIXRD) and residual stress varying layer thickness between h = 2.5 and 30 nm.
measurement at 45 kV and 40 mA. The conventional Periodic boundary conditions were applied to the X
Bragg–Brentano configuration was selected to keep and Y directions while the Z direction was set free. The
the substrate peak intensity proportionally low, com- coherent Al/Ti {111}/{0001} interfaces were created by
pared to the constituent phases of the thin films, and scaling the lattice constant of Ti to match that of Al,
a grazing incident angle (α) of 3° was used, giving while the incoherent Al/Ti {111}/{0001} interfaces were
an estimated penetration depth of ~ 1.0 μm that is created without any scaling. For the latter, interfacial
much larger than h of all NMMs. According to ­sin2Ψ dislocations were generated after a relaxation by using
method, the diffraction planes of Al (220) and Ti (101 3) conjugate gradient energy minimization at 0 K. Prior

Table 1  Summary of
Sample IDs Hardness Reduced elastic modulus Individual layer Total film
mechanical properties,
(H, GPa) (Er, GPa) thickness thickness
individual layer thickness (h)
(h, nm) (nm)
and total film thickness layer
of monolithic Al, Ti thin Al/Ti 2.5 4.94 ± 0.44 120.14 ± 7.56 2.5 ± 0.3 930
films and Al/Ti NMMs Al/Ti 5.2 4.85 ± 0.28 117.57 ± 2.91 5.2 ± 0.6 820
Al/Ti 11.4 4.16 ± 0.49 112.89 ± 7.97 11.4 ± 1.1 1566
Al/Ti 30 3.13 ± 0.43 109.99 ± 7.22 30 ± 1.5 1800
Al/Ti 36 2.87 ± 0.43 108.99 ± 8.83 36 ± 1.9 2650
Al/Ti 52 2.45 ± 0.29 106.72 ± 9.17 52 ± 2.4 3000
Al 0.67 ± 0.17 87.2 ± 14.90 N/A 2000
Ti 4.62 ± 0.62 119.15 ± 11.13 N/A 1200

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

is set to be 40 nm, and K is 10 eV Å.−3. An isother-


mal–isobaric ensemble (NPT) was maintained during
the indentation process (with pressure P = 0 along the
X and Y directions and temperature T = 300 K). The
hardness of each Al/Ti NNM was estimated by com-
puting the mean contact pressure ­Pm at the yield point
as [29]

Pm = p/A (2)
Figure 1  a Schematic nanoindentation model set-up of Al/Ti where p is the total load applied by the indenter and A
multilayers, b MD simulation of interface structures with coher- is the projected contact area. A can be approximated as
ent and incoherent, and green atoms represent Al, orange atoms A = a*LY where a was directly measured from the snap-
represent Ti.
shot of the indented sample as shown in Fig. 1a. Post-
visualization and analysis of the simulation results
to nanoindentation, all models were equilibrated at were mainly performed using OVITO [30] with the
300 K under the isothermal–isobaric ensemble (NPT) common neighbor analysis [31] and local atomic shear
and zero pressure for 25 ps. After that, a thin slab strain [32].
(~ 0.5 nm thick) at the bottom of the model was fixed
and the indentation was performed by following the
same protocol as described in [28], i.e., by moving a
Results and discussion
cylindrical virtual indenter at a constant velocity of
5 m/s along the Z direction. During this process, a
Microstructure and residual stress
repulsive force was exerted to the Al/Ti NNMs accord-
ing to the following equation:
Figure 2 shows the SEM images of as-deposited Al/
p(r) = K(r − R) 2
(1) Ti multilayers with nanoscale surface nodules, rep-
resenting the formation of ultrafine microstructures.
where p(r) is the repulsive force, R is the radius of the The arithmetic average surface roughness ­(Ra) and
indenter, r is the distance from the atoms in the Al/Ti nodule diameter of all Al/Ti NMMs were summa-
NNMs to the center axis of the cylindrical indenter, rized in Fig. 3. As a comparison, the surface proper-
and K is the specified force constant. In this work, R ties of monolithic Al and Ti films were also displayed.

Figure 2  SEM images of the surface of as-deposited Al/Ti NMMs with different individual layer thickness.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

Samples with thicker h can be found with coarser nod- other hand, the Al layers developed mostly compres-
ules and rougher surface, despite the slight deviation sive stresses, where the highest compressive stress was
at h = 2.5 nm, as reported previously [33]. It should be observed at h = 2.5 nm. It is generally believed that
noted that the nodule size often corresponds to the internal tensile stress will decrease the hardness while
in-plane grain size (d), which affects the mechanical compressive stress does not have a significant effect on
properties in polycrystalline multilayered thin films hardness [18]. However, at present, it is a challenge
[34]. However, in this work, since d is much larger to separate the effects of both tensile and compres-
than h, thus the smaller of the two (i.e., h) would be the sive stress originated from the individual layers on
dominant factor on the strength. Figure 4a shows the the overall mechanical properties of the multilayers.
GIXRD patterns of all multilayers and Fig. 4b plots the
residual stress measured from Al (220) and Ti (101 3) Length scale‑dependent hardness
diffractions. Regardless of the size of h, tensile stress is
built up in Ti layers, which is consistent with previous The hardness and reduced elastic modulus of mon-
results under similar deposition conditions [20]. The olithic Al, Ti films and Al/Ti multilayered samples
magnitude of the residual tensile stress in Ti layers are summarized in Table 1. The hardness gradually
reaches above 2 GPa when h is below ~ 10 nm. On the increased as h decreased, reaching the peak strength of
~ 4.94 GPa at h = 2.5 nm, indicating the size-dependent
deformation behaviors. Figure 5a shows the hardness
versus the inverse square root of individual layer
thickness (h−1/2) for Al/Ti multilayer and other typi-
cal NMMs including FCC or HCP phase [6, 35, 36].
Among all samples plotted, Al/Ti multilayers exhib-
its the highest hardness at small h. At thick h, the
hardness is in agreement well with the prediction of
Hall–Petch model, which is defined as:

H = H0 + kh−1∕2 (3)

where H is the material hardness, ­H0 (~ 0.32 GPa) is


the intercept point with y axis on H versus ­h−1/2 plot, k
(~ 15.23 GPa (nm)1/2) is the Hall–Petch slope obtained
from fitting the linear section of data points (i.e., the
red dashed line in Fig. 5a). As discussed in the intro-
Figure 3  Summary of grain diameter (square symbol) and
duction, the Hall–Petch model will severely overes-
arithmetic average surface roughness ­(Ra) measured from SEM
timate strength when h is further decreased due to
images of all as-deposited Al/Ti NMMs and pure Ti and Al thin
films. the insufficient space of layers for the propagation of

Figure 4  a GIXRD patterns


and b measured residual
stress of Al and Ti layers
of all as-deposited Al/Ti
NMMs.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

Figure 5  a Hardness as a function of individual layer thickness Ti NMMs. The scatted data corresponds to experimental meas-
for Al/Ti and other typical NMMs including FCC or HCP phases. urements and the solid line corresponds to predications by CLS
b Hardness as a function of individual layer thickness of all Al/ model.

dislocations. When h is around a few tens of nanom- TEM characterization of indented structure
eters, the plastic flow will be constrained to one layer
and the single dislocation glides tend to propagate To characterize their deformed microstructure caused
with the layer in the form of hairpin [11]. In this by nanoindentation, representative HRTEM cross-sec-
regime, CLS model is more appropriate to estimate the tional images of the h = 30 nm and h = 2.5 nm samples
shear stress required to cause plastic deformation as: were shown in Fig. 6. Typical chemically modulated
structures and the layer thickness refinement can
𝛼h�
[ ]
𝜇b ( 4 − 𝜈 )
τcls = M ⋅ ⋅ ⋅ ln
f C
− + (4) be clearly seen around the deformed area when h is
8𝜋h� 1−𝜈 b h 𝜆 thick (Fig. 6b) with incoherent interfaces. While the
and more coherent interfaces are observed at h = 2.5 nm in
Fig. 6a, indicating the phase transformation from HCP
𝜇b to FCC in the Ti layer at small h, consistent with the
C= (5)
(1 − 𝜈) results in Fig. 4a and our previous work [24]. How-
ever, it should also be pointed out that earlier work
[37] on Al/Ti multilayers showed that fcc Ti is only
where M (M = 3.1) is the Taylor factor, μ observed in ion-milled TEM foils of Al/Ti multilayer
(μ = 44 GPa) is the shear modulus, b (b = 0.17 nm) is the thin films and not in X-ray diffraction. The impact of
Burgers vector, h′ is the layer thickness measured par- impurities (e.g., hydrogen) [38] during TEM thin-foil
allel to the slip plane, ν (ν = 0.32) is the Poissonʼs ratio, milling [39] could be one of the reasons. Other factors,
f is the interface stress, α is the core cut-off param- including the effects of stacking sequence change and
eter, and λ is the spacing of the interface which can energy tradeoff between the bulk elastic strain energy
be found by fitting the experimental results from the and surface free energy, on the Ti layer phase transfor-
current research. The tendency of calculated hardness mation has been discussed in our previous study [24]
by CLS model is shown in Fig. 5b, which is consistent and by others [5, 40]. Next, using molecular dynam-
with experimental results, confirming that this model ics simulations, we further show that such differences
is effective in the current length scales. Notably, an in interface coherency further affects the deformation
overshoot of the CLS predicted hardness is found at behavior, even under the same h.
around 2–5 nm, suggesting that a different deforma-
tion mechanism is at play, where strength depends
on the interface coherency and the interface barrier
strength, instead of layer thickness.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

Figure 6  High resolution


bright field TEM images
after nanoindentation of a Al/
Ti at h = 2.5 nm and b Al/
Ti at h = 30 nm. Inserted in
a is zoomed-out of Al/Ti at
h = 2.5 nm after nanoindenta-
tion.

Effects of interface coherency and layer incoherent type interfaces is considered to be larger
thickness on deformation mechanisms than the coherent interfaces when h is thick, causing
higher hardness. When the layer thickness is relatively
To understand the effects of h and interface coherency small (i.e., h = 2.5 nm), the underlying plastic defor-
on the deformation behaviors of Al/Ti NMMs at the mation becomes dominated by dislocation slip across
atomic scale, MD simulated indentation model was set the interfaces. Unlike the coherent interface, interfacial
up by two types of interface microstructure: coherent dislocation can be observed in the Al/Ti interfaces due
and incoherent, which depends on the atomic match- to the high lattice mismatch of Al and Ti (Fig. 9b). At
ing degree across the interface [41], as shown in Fig. 1. this scale, not only is the strength of Al layer increased,
In comparison, the load-penetration depth curves of resulting in obvious co-deformation between Al and Ti
all Al/Ti NMMs samples with incoherent interface layers, but also a HCP to FCC phase transformation of
(Fig. 7a) and coherent interface (Fig. 7b) after indenta- Ti layer is observed (circled areas in Fig. 9c). In Fig. 9a,
tion were presented. It can be seen that the load and the dislocation core at the stress concentration area
penetration depth at which yielding occurrence (i.e., in incoherent interface was observed at the penetra-
the first peak) are larger in the structures with inco- tion depth 3.75 nm and rapidly propagates towards
herent interfaces than the coherent ones, except at deep layers when penetration depth reached 3.8 nm
h = 2.5 nm. These results indicate that the incoherent (Fig. 9b), accompanied with interface rotation. How-
interface state causes higher hardness than the coher- ever, as for coherent interface (Fig. 9c), the network of
ent interface when h exceeds 4 nm, but the coherent dislocation was only localized in the top few layers,
interface leads higher hardness at h = 2.5 nm (Fig. 7c), suggesting its higher plastic deformation resistance
fully consistent with the experimental results. than the incoherent interface.
The subsurface residual strain and dislocation
movements when first yielding occurred at h = 15 nm
and h = 2.5 nm, representing the typical contrast
between coherent and incoherent interfaces, were Conclusions
presented in Figs. 8 and 9. At h = 15 nm, regardless
of incoherent and coherent interfaces, the plastic flow In summary, the hardness, residual stress, and defor-
will be constrained in the soft Al layer (Fig. 8), simi- mation mechanism of Al/Ti NMMs with alternating
lar to the previously reported in Cu/Zr multilayer FCC/HCP structure are evaluated by nanoindentation
[42]. The dislocations are absorbed and dispersed at experiments, advanced materials characterization,
the interfaces, resulting in layer bending as a result and molecular dynamics simulations. In general, a
of that applied stress, which is smaller than the Al/ tensile stress develops in the Ti layer and compres-
Ti interface barrier stress. In addition, due to the dis- sive stress in the Al layers, with increasing magnitude
continuity in dislocation slip, the barrier stress of at smaller h. At the same time, hardness increases as
the layer thickness decreases, reaching a peak strength

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

Figure 7  Nanoindenta-
tion force-penetration depth
curves of a incoherent Al/
Ti and b coherent Al/Ti with
different h, and c summary
of simulated hardness as a
function of h under different
interface structures.

Figure 8  MD simulation


results of von Mises atomic
strain and dynamic response
after nanoindentation of a
incoherent Al/Ti at h = 15 nm
and b coherent Al/Ti at
h = 15 nm.

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

Figure 9  MD simulation results of von Mises atomic strain row images represent the dislocation activities via the bottom row
and dynamic response after nanoindentation of incoherent Al/Ti images represent the corresponding phases (green: aluminum
2.5 nm at penetration depth of a 3.75 nm and b 3.8 nm, and c FCC, orange: titanium HCP).
coherent Al/Ti 2.5 nm at penetration depth of 2.9 nm. The top

of ~ 4.94 GPa at 2.5 nm layer thickness, owing to its Author contributions


length-scale-dependent deformation characteristics. In
thick h regime, the predicted strength by CLS model W W: Data curation, Visualization, Investigation,
is well consistent with the experimental results and Writing-Original Draft. SI: Data curation, Visualiza-
MD simulation results reveal that the interface bar- tion, Investigation. HM: Data curation, Investigation.
rier strength of incoherent interface is larger than DC: Software, Investigation, Writing-Reviewing and
that of coherent interfaces. However, at h = 2.5 nm, Editing. WC: Writing-Reviewing and Editing, Fund-
coherent interfaces will cause higher hardness and ing acquisition.
a deviation from CLS model. Molecular dynamics
simulation shows that at such small h, the coherent
interfaces becomes more effective barriers to disloca- Declarations
tion transfer and the first dislocation nucleation occurs
at much smaller depths than those of the incoherent Conflict of interest The authors declare that there is
counterparts. no conflict of interest.

Ethical approval Not applicable.

Data and code availability

The data that supports the findings of this study References:


are available upon request from the corresponding
authors. [ 1] Xie TT, Mao SD, Yu C, Wang SJ, Song ZL (2012) Struc-
ture, corrosion, and hardness properties of Ti/Al multilay-
ers coated on NdFeB by magnetron sputtering. Vacuum
Acknowledgements 86(10):1583–1588
[ 2] Creus J, Top EH, Savall C, Refait P, Ducros C, Sanchette
This research was financially supported by the F (2008) Mechanical and corrosion properties of dc mag-
US National Science Foundation under Grant netron sputtered Al/Cr multilayers. Surf Coat Technol
CMMI-1855651. 202(16):4047–4055

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

[ 3] Wang J, Misra A (2023) Plastic homogeneity in nanoscale [18] Huber N, Heerens J (2008) On the effect of a general resid-
heterostructured binary and multicomponent metallic ual stress state on indentation and hardness testing. Acta
eutectics: an overview. Curr Opin Solid State Mater Sci Mater 56(20):6205–6213
27(1):101055 [19] Goudeau P, Badawi KF, Naudon A, Gladyszewski G (1993)
[ 4] Misra A, Göken M, Mara NA, Beyerlein IJ (2021) Hierar- Determination of the residual-stress tensor in Cu/W multi-
chical and heterogeneous multiphase metallic nanomateri- layers by X-ray-diffraction. Appl Phys Lett 62(3):246–248
als and laminates. MRS Bull 46:236–243 [20] Ljungcrantz H, Hultman L, Sundgren JE, Johansson
[ 5] Zhang Y, Xue S, Li Q, Li J, Ding J, Niu T, Su R, Wang S, Kristensen N, Schweitz JÅ, Shute C (1993) Residual
H, Zhang X (2019) Size dependent strengthening in high stresses and fracture properties of magnetron sputtered Ti
strength nanotwinned Al/Ti multilayers. Acta Mater films on Si microelements. J Vac Sci Technol A: Vac Surf
175:466–476 Films 11(3):543–553
[ 6] Lu Y, Kotoka R, Ligda J, Cao B, Yarmolenko S, Schuster [21] Stafford GR, Kongstein OE, Haarberg GM (2006)
B, Wei Q (2014) The microstructure and mechanical behav- In situ stress measurements during aluminum deposition
ior of Mg/Ti multilayers as a function of individual layer from ­A lCl 3-EtMeImCl ionic liquid. J Electrochem Soc
thickness. Acta Mater 63:216–231 153(4):C207
[ 7] Misra A, Hirth J, Hoagland R (2005) Length-scale-depend- [22] Gu X, Zhang Z, Bartosik M, Mayrhofer PH, Duan H (2017)
ent deformation mechanisms in incoherent metallic multi- Dislocation densities and alternating strain fields in CrN/
layered composites. Acta Mater 53(18):4817–4824 AlN nanolayers. Thin Solid Films 638:189–200
[ 8] Wu S, Hou Z, Zhang J, Wang Y, Wu K, Liu G, Sun J (2019) [23] Todt J, Krywka C, Zhang ZL, Mayrhofer PH, Keckes J,
Understanding the strength softening of Zr/Mo and Zr/Ti Bartosik M (2020) Indentation response of a superlattice
nanostructured multilayers. Scr Mater 172:61–65 thin film revealed by in-situ scanning X-ray nanodiffrac-
[ 9] Misra A, Verdier M, Kung H, Embury JD, Hirth JP (1999) tion. Acta Mater 195:425–432
Deformation mechanism maps for polycrystalline metallic [24] Izadi S, Mraied H, Cai W (2015) Tribological and mechani-
multiplayers. Scr Mater 41(9):973–979 cal behavior of nanostructured Al/Ti multilayers. Surf Coat
[10] Misra A, Kung H (2001) Deformation behavior of Technol 275:374–383
nanostructured metallic multilayers. Adv Eng Mater [25] Wang W, Wang K, Zhang Z, Chen J, Mou T, Michel FM,
3(4):217–222 Xin H, Cai W (2021) Ultrahigh tribocorrosion resistance of
[11] Wang J, Misra A (2011) An overview of interface-dom- metals enabled by nano-layering. Acta Mater 206:116609
inated deformation mechanisms in metallic multilayers. [26] Thompson AP, Aktulga HM, Berger R, Bolintineanu DS,
Curr Opin Solid State Mater sci 15(1):20–28 Brown WM, Crozier PS, in’t Veld PJ, Kohlmeyer A, Moore
[12] Misra A, Kung H, Embury J (2004) Preface to the view- SG, Nguyen TD, Shan R, Stevens MJ, Tranchida J, Trott C,
point set on: deformation and stability of nanoscale metal- Plimpton SJ (2022) LAMMPS-a flexible simulation tool for
lic multilayers. Elsevier, The Netherlands particle-based materials modeling at the atomic, meso, and
[13] Thomas O, Gergaud P, Labat S, Barrallier L, Charaï A, continuum scales. Comput Phys Commun 271:108171
Alfonso C, Gilles B, Marty A (1996) Residual stresses in [27] Zope RR, Mishin Y (2003) Interatomic potentials for
metallic multilayers. J Phys IV 6(C7):C7-125 atomistic simulations of the Ti-Al system. Phys Rev B
[14] Labat S, Gergaud P, Thomas O, Gilles B, Marty A (2000) 68(2):024102
Interdependence of elastic strain and segregation in metal- [28] Deng C, Schuh CA (2012) Atomistic mechanisms of cyclic
lic multilayers: an X-ray diffraction study of (111) Au/Ni hardening in metallic glass. Appl Phys Lett 100(25)
multilayers. J Appl Phys 87(3):1172–1181 [29] Dupont V, Sansoz F (2009) Molecular dynamics study of
[15] Berger S, Spaepen F (1995) The Ag/Cu interface stress. crystal plasticity during nanoindentation in Ni nanowires.
Nanostruct Mater 6(1–4):201–204 J Mater Res 24(3):948–956
[16] Bain JA, Chyung LJ, Brennan S, Clemens BM (1991) Elas- [30] Stukowski A (2010) Visualization and analysis of atomistic
tic strains and coherency stresses in Mo/Ni multilayers. simulation data with OVITO–the open visualization tool.
Phys Rev B 44(3):1184–1192 Model Simul Mater Sci Eng 18(1):015012
[17] Chang RC, Chen FY, Chuang CT, Tung YC (2010) [31] Honeycutt JD, Andersen HC (1987) Molecular dynamics
Residual stresses of sputtering titanium thin films at study of melting and freezing of small Lennard–Jones clus-
various substrate temperatures. J Nanosci Nanotechnol ters. J Phys Chem 91(19):4950–4963
10(7):4562–4567

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


J Mater Sci

[32] Shimizu F, Ogata S, Li J (2007) Theory of shear banding [40] Zhang Y, Li Q, Gong M, Xue S, Ding J, Li J, Cho J, Niu T,
in metallic glasses and molecular dynamics calculations. Su R, Richter N (2020) Deformation behavior and phase
Mater Trans 48:2923–2927 transformation of nanotwinned Al/Ti multilayers. Appl
[33] Pobedinskas P, Bolsée J-C, Dexters W, Ruttens B, Mortet Surf Sci 527:146776
V, D’Haen J, Manca JV, Haenen K (2012) Thickness [41] An M, Deng Q, Su M, Song H, Li Y (2017) Dependence of
dependent residual stress in sputtered AlN thin films. Thin deformation mechanisms on layer spacing in multilayered
Solid Films 522:180–185 Ti/Al composite. Mater Sci Eng: A 684:491–499
[34] Misra A, Verdier M, Lu Y, Kung H, Mitchell T, Nastasi M, [42] Zhang J, Lei S, Liu Y, Niu J, Chen Y, Liu G, Zhang
Embury J (1998) Structure and mechanical properties of X, Sun J (2012) Length scale-dependent deformation
Cu-X (X = Nb, Cr, Ni) nanolayered composites. Scr Mater behavior of nanolayered Cu/Zr micropillars. Acta Mater
39(4–5):555–560 60(4):1610–1622
[35] Fu E, Li N, Misra A, Hoagland R, Wang H, Zhang X (2008)
Mechanical properties of sputtered Cu/V and Al/Nb multi- Publisher’s Note Springer Nature remains neutral with
layer films. Mater Sci Eng: A 493(1–2):283–287 regard to jurisdictional claims in published maps and
[36] Ham B, Zhang X (2011) High strength Mg/Nb nanolayer institutional affiliations.
composites. Mater Sci Eng: A 528(4–5):2028–2033
Springer Nature or its licensor (e.g. a society or other partner)
[37] Banerjee R, Dregia SA, Fraser HL (1999) Stability of fcc
holds exclusive rights to this article under a publishing
titanium in titanium/aluminum multilayers. Acta Mater
agreement with the author(s) or other rightsholder(s);
47(15–16):4225–4231
author self-archiving of the accepted manuscript version of
[38] Banerjee R, Zhang X-D, Dregia S, Fraser H (1999) Phase
this article is solely governed by the terms of such publishing
stability in Al/Ti multilayers. Acta Mater 47(4):1153–1161
agreement and applicable law.
[39] Bonevich J, Van Heerden D, Josell D (1999) Face-centered-
cubic titanium: an artifact in titanium/aluminum multilay-
ers. J Mater Res 14:1977–1981

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:

1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at

onlineservice@springernature.com

You might also like