Quantitative Nanoscale Magnetic Study of Isolated Diameter

Sep 28, 2016 - Rotate animation of the simulated magnetization distribution of Figure 6a ...... by a Stray Field from NiFe on a Co/Ni Nanowire J. Kore...
0 downloads 0 Views 2MB Size
Quantitative Nanoscale Magnetic Study of Isolated Diameter-Modulated FeCoCu Nanowires Luis Alfredo Rodríguez,*,‡ Cristina Bran,† David Reyes,‡ Eider Berganza,† Manuel Vázquez,† Christophe Gatel,‡ Etienne Snoeck,‡ and Agustina Asenjo† ‡

CEMES-CNRS 29, rue Jeanne Marvig, B.P. 94347, F-31055 Toulouse Cedex, France Instituto de Ciencia de Materiales de Madrid (ICMM-CSIC), Cantoblanco, Madrid 28049 Spain



S Supporting Information *

ABSTRACT: The comprehension of the magnetic configuration in FeCoCu nanowires with a diameter-modulated cylindrical geometry will allow controlling the domain wall motion in this low-dimensional system under the application of magnetic fields and/or the injection of current pulses. Here we perform a quantitative magnetic characterization of isolated diameter-modulated FeCoCu nanowires by combining nanoscale magnetic characterization techniques such as electron holography, magnetic force microscopy, and micromagnetic simulations. Local reconstructions of the magnetic distribution show the diameter-modulated geometry of the wires induces the formation of vortex-like structures and magnetic charges in the regions where the diameter is varied. Vortex-like structures modify the axial alignment of the magnetization in large-diameter segments. Moreover, the magnetic charges control the demagnetizing field distribution, promoting a flux-closure stray field configuration around large-diameter segments and keeping the demagnetizing field parallel to the NW’s magnetization around small diameter segments. The detailed description of the remanent state in diameter-modulated cylindrical FeCoCu nanowires allows us to provide a clear explanation of the origin of bright and dark contrast observed in magnetic force microscopy images, which have the same feature of magnetic domain walls. This work establishes the primary knowledge required for future magnetization reversal studies with the aim of searching efficient modulated geometries that allow an optimum and controlled domain wall propagation. KEYWORDS: magnetic nanowires, electron holography, micromagnetic simulations, spin configuration, magnetic charges, magnetic imaging

N

diameter of cylindrical NWs is expected to be efficient to control the DW propagation. Micromagnetic and Monte Carlo simulations18−21 have shown that such modulation of the NW diameter can indeed pin DWs, and experimental studies performed by magnetic force microscopy (MFM), X-ray magnetic circular dichroism, and magneto-optic Kerr effect22−24 have confirmed this DW pinning behavior. In parallel, recent improvements on modern preparation methods that follow electrochemical route allow fabricating simultaneously a great number of cylindrical NWs with length of tens of microns and the possibility to tune the diameter-modulation (DM) geometry of the wires.25−28 Thus, the combination of a high

owadays there is an increasing interest for searching novel magnetic nano-objects, allowing the control of individual magnetic domain walls (DWs) motion. One-dimensional (1D) nanostructures such as ferromagnetic self-standing nanowires (NWs) or patterned planar are used to achieve a precise DW manipulation either by the application of magnetic fields or by the injection of electric current. Magnetic 1D nanostructures are therefore of large interest for possible application in the new-generation spintronic devices for information storage technologies, permanent magnets, logic systems, sensors, and biomedical applications.1−8 Beside the intrinsic DW pinning behavior,9,10 one of the most promising routes to manipulate DWs in planar nanostrips is by patterning lateral defects like notches or antinotches that act as DW nucleation as well as extrinsic pinning and depinning centers of DWs.11−17 Inspired by the same idea, modulation in the © 2016 American Chemical Society

Received: August 15, 2016 Accepted: September 28, 2016 Published: September 28, 2016 9669

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

www.acsnano.org

Article

ACS Nano

Figure 1. Bright-field TEM images of two representative DM FeCoCu NWs of (a) 13 and (b) 4.7 μm in length. The NW diameters and the segment lengths are measured in the high-magnification TEM images corresponding to the areas enclosed in (a) by (c) yellow and (d) red boxes. (c,d) Illustration of the progressive change of the segments length.

Figure 2. HRTEM images taken in two zones along the one DM FeCoCu NW: (a) SD and (b) LD segments. As inset of each TEM image, the diffraction rings in the FFTs demonstrate the polycrystalline structure of the NWs.

characterization of the magnetic induction distribution within and around isolated cylindrical NWs.31,32 However, the measured signal corresponds to an integration of the all inplane components of the magnetic induction along the electron path, and, if a 3D tomographic reconstruction is not carried out,33−35 EH is not thus sufficient to obtain a full description of the local configuration of the magnetization. Nevertheless, several works have demonstrated that a quantitative combination of EH and micromagnetic simulations will be able to elucidate the 3D magnetic states in ferromagnetic NWs.31,32 In this work, we have exploited the potential of EH and micromagnetic simulations to perform a high-resolution quantitative magnetic characterization of cylindrical FeCoCu NWs with a DM geometry. The possibility to reconstruct the magnetic induction field distribution both inside (magnetization) and outside (stray fields) of the NW allows a deeper insight on the optimal geometry characteristics to achieve a controlled DW propagation.

DW propagation velocity favored by the absence of a Walker breakdown phenomenon29 with the most efficient DM geometry to achieve a precise DW manipulation makes cylindrical NWs more powerful systems compared to their planar counterpart. However, the achievement of a controlled and reproducible DW handling requires a preliminary exhaustive study focused on understanding how the remanent local magnetic configuration is altered by changes of the lateral geometry in individual and isolate NWs, free of magnetic interaction among neighboring nanowires. In a previous work, highly sensitive MFM measurements combined with micromagnetic simulations allowed a preliminary approach of the spin configuration of much thinner individual DM FeCoCu NWs. 30 At remanence, bright and dark contrast appeared along the wire associated with the increase of the stray field in the transition zone where the diameter changes. The nontrivial interpretation performed in ref 30 can be substantially improved by performing more resolved magnetic imaging techniques such as electron holography (EH). This transmission electron microscopy (TEM)-based method allows magnetic mapping at the nanoscale with high sensitivity and provides quantitative local analysis by volume which permits a more detailed

RESULTS AND DISCUSSION Intermediate-magnification TEM images displayed in Figure 1 illustrate the morphology of two representative DM NWs 9670

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano

Figure 3. (a) Topographic and (b) MFM images for a DM FeCoCu NW of intermediate length.

technique is sensitive to the magnetostatic interaction between the magnetized probe and the out-of-plane component of the stray field; the high-contrast spots in the magnetic image correspond to the regions with intense stray field parallel or antiparallel to the MFM tip and perpendicular to the NW surface. Head-to-head and tail-to-tail DWs could produce this kind of magnetic contrast because the progressive rotation of the magnetization within the wall creates a magnetic volume charge where the stray field is coming in and out of the wire. In addition, the formation of a series of DWs along the wire could also create a periodic variation of bright and dark spots because the dipolar effect of each small magnetic domain will induce magnetic poles of opposite sign in the extremities of each domain. On the other hand, bright/dark spots can be also produced by a complex magnetization configuration caused by the progressive variation of the diameter. In ref 30, the formation of curling states in the DM transition was associated with such high-contrast spots. Thus, the limitation of MFM technique to map only the stray field does not allow us to distinguish what type of magnetization structure (either DWs, curling states or both) causes such magnetic signal, so complementary nanoscale magnetic characterizations able to resolve the local magnetization distribution such as EH and micromagnetic simulation are required. A quantitative analysis of the magnetic distribution in DM FeCoCu NWs was carried out by EH experiments. This technique provides information on the in-plane component of the magnetic induction by means of the retrieved phase shift of an electron wave interacting with a ferromagnetic material (see Supporting Information). The projection of the in-plane components of the magnetic induction Bxy = (Bx, By) can be obtained by differentiating the magnetic contribution of the phase shift, φM:

prepared by electrodeposition. Among the numerous NWs deposited on the TEM carbon grid, the longest ones (similar to the NW reported in Figure 1a) are about 13 μm long, although most of the NWs look similar to the wire in Figure 1b with intermediate lengths between 3 and 8 μm. These 1D nanostructures are formed by alternating segments of small diameter (SD, around 100 nm) and large diameter (LD, around 140−144 nm) with a smooth diameter-variation transition. If we move from the right to left side along the NW in Figure 1a, we see that the length of the SD and LD segments is gradually reduced, with a more pronounced variation in the LD segments. The LD segments are around 1000 nm long at the right end and nearly disappear at the left end. In the case of SD segments, their lengths are reduced from 430 to 300 nm (see Figure 1c,d corresponding to the zoom of the marked areas in Figure 1a). This geometrical variation previously observed in DM Co NWs, also fabricated by electroplating in AAO templates, is attributed to plastic deformation and mechanical instabilities on the aluminum substrate. 36 From TEM observations (see Figure 1a,b) we deduce that NWs of intermediate lengths are actually longer pieces that break during the extraction from the template. Microstructural analysis of the wires was performed in highresolution TEM (HRTEM) mode, as reported in Figure 2. These HRTEM images, combined with electron diffraction experiments (not shown), do not reveal a clear single-crystal structure of any texture over a long distance. Instead, the wires present a polycrystalline structure formed by nanocrystals of different sizes in the range of few nanometers. The diffraction rings observed in the fast Fourier transforms (FFTs) of the HRTEM images certify the polycrystalline character of the wires with randomly oriented nanograins. In addition, the surface of the NWs is covered by a 2−3 nm-thick amorphous layer. This superficial amorphous layer formed by a surface oxidation process has been previously observed in FeCoCu NWs.30,37 The first analysis of the remanent magnetic state of the DM FeCoCu NWs was carried out by MFM measurements. Figure 3 displays topographic and magnetic images taken in an intermediate-length isolated NW with an as-grown magnetic state (no magnetic field has been applied). Similar to that observed in the ref 30, the MFM image of the wire shows extended regions with a very weak magnetic contrast separated by high-contrast bright and dark spots located at the ends of the wire and in the regions where the diameter is varied. As MFM

∂φM(x , y) ∂x

=

e ℏ



∫−∞ By(x , y , z)dz

(1)

and ∂φM(x , y) ∂y

=−

e ℏ



∫−∞ Bx (x , y , z)dz

(2)

where e is the electron charge, and ℏ is the reduced Planck constant. As EH only provides information on the z-axis integration of the magnetic induction, a precise description of the 3D local spin configuration was supported by micro9671

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano

Figure 4. (a−d) Experimental and (e−h) simulation analysis of the magnetic phase shift of the electron produced by the magnetic induction of an isolated DM FeCoCu NW: (a, e) Morphology of the projected nanowire image and (b, f) magnetic phase shift image reconstructed for a section of the NW. (c, g) and (d, h) Magnetic flux images reconstructed from the magnetic phase shift images using an amplifier factor of n = 2 and 8, respectively. (i) Magnetic phase shift profiles extracted from (b) (red and blue arrows) and (f) (black and gray arrows), respectively. Color in the magnetic flux images indicates only the magnetic induction direction according to the inserted color wheel, and the white arrows in (d) help to visualize the local orientation of the in-plane projected magnetic induction inside and around the NW.

experimental and simulated results. The optimization of such experimental and simulation concordance was achieved by adjusting some of the magnetic parameters used in the micromagnetic simulation such as the saturation magnetization MS and the exchange constant A (anisotropy constant K was neglected due to the polycrystalline character of the wires formed by randomly oriented nanocrystals) and performing a quantitative comparison between experimental and simulated magnetic phase shift images through profiles traced perpendicular to the NW axis. Figure 4i shows experimental and simulated magnetic phase shift profiles taken in the middle of a SD and LD segment: The best matching was achieved by simulating the remanent magnetic state with a saturation magnetization value of μ0MS = 1060 kA/m (∼1.33 T) and using a value of A = 26 × 10−12 J/m which corresponds to that found for the Fe30Co70 alloy.38,39 In this way, we indirectly determine the saturation magnetization of the DM FeCoCu NWs, and this value is much lower than those measured through a vibrating sample magnetometer in FeCoCu NW arrays either in an as-prepared condition (2.0 T) or after annealing the sample at 500 °C (1.7 T).40 Magnetic phase shift image of Figure 4b presents a strong color variation inside the NW, along its radial axis (y-axis). According to the eq 2, this color (or phase shift) variation indicates that the projected magnetic induction is mainly oriented along the NW axis. This fact can be visualized in the magnetic flux image of Figure 4c where the direction of the flux

magnetic simulations. The simulated magnetic distribution at remanence of a DM FeCoCu NW was obtained on a wire of 2.6 μm length consisting of only four repetitions of SD and LD segments to reduce the computational time. The diameter was varied between 100 and 140 nm, and the lengths of each type of segment were kept constant with values of 300 nm (SD segment) and 250 nm (LD segment) with a linear diametervariation transition of 50 nm length. In addition, we have built the NW with ends of different diameter in order to evaluate their influence on the local magnetization orientation. The geometrical parameters such as the size and shape of the DM NW, as well as the size of the simulation universe, are schematically illustrated in Figure 8. Figure 4 displays EH images of the amplitude and magnetic phase shift at the remanent state in an intermediate region of the NW reported in Figure 1. In addition, two magnetic flux images, Φ(x, y), are also presented in Figure 4 which are reconstructed by the application of a cosine function on an amplified magnetic phase shift image of Figure 4b [i.e., Φ(x, y) ∼ cos(n·φM(x, y)); n is an amplifier factor]. As observed in the amplitude image of the Figure 4a, the studied area contains a part of the wire composed by four repetitions of SD and LD segments with lengths ranging between 300 and 270 nm and 430 and 300 nm, respectively. Similar magnetic phase shift and magnetic flux were reconstructed from the micromagnetic simulations of an intermediate section of the NW of Figure 8. An excellent agreement has been obtained between the 9672

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano

Figure 5. Magnetic flux images reconstructed around two NW ends with (a) small and (b) large diameters. The correspondent calculated magnetic flux images for a (c) SD and (d) LD NW end. The colormap represents the magnetic induction direction according to the color wheel located beside (d). Green contour lines and gray regions in (a) and (b) are artificially added to delineate the NW edges and to highlight the carbon TEM support, respectively.

the wire in the transition regions where the diameter changes. This interesting stray field configuration suggests that each diameter-variation crossover acts as a positive or negative magnetic charge that controls the local configuration of the demagnetizing field. According to the colormap used in the magnetic flux images, the positive and negative character of the magnetic charges changes alternately along the NW, and the sign is correlated with the flux direction of the magnetic induction inside the wire: if the field flows from a LD (SD) segment to a SD (LD) segment, the crossover between them behaves as a positive (negative) charge. As the magnetic induction inside the wire flows in the same direction, the remanent magnetic state of the DM NWs induces a sequence of alternately positive and negative magnetic charges. Making a local analysis, we found that these magnetic charge arrangements produce a flux-closure configuration of the demagnetizing field around LD segments, and they force the stray field flow parallel to the magnetization around SD segments. This remanent magnetic distribution resembles that observed in the segmented Co/Ni NW, where a compositional variation creates a modulation of the magnetization amplitude because the magnetic moment of the Co is 3 times larger than Ni. Thus, this magnetization modulation promotes the formation of magnetic poles in the Co−Ni interfaces that act similar to those of the DM FeCoCu NW.41,42 To describe the overall remanent magnetic state of DM FeCoCu NWs, magnetic flux images around the ends of the wires were also reconstructed. Figure 5 displays magnetic images of the two possible types of ends (SD and LD end). Additionally, magnetic flux images around the ends of the simulated DM NW were also represented (see Figure 5c,d). As the shape of the ends in the real NWs is more irregular than that of the simulated one, the configuration of the magnetic flux lines is not perfectly comparable, but they have common

lines into the nanowires shows a longitudinal alignment of the magnetization. However, we detect a small difference in the configuration of the flux lines: inside the SD segments, the lines are perfectly parallel to the NW axis, and inside the LD segments, the lines are bended as they approach to the lateral sides (a zoom of this area is shown in the Supporting Information). This effect suggests that the longitudinal alignment of the magnetization could be modified in LD segments. A magnetization parallel to the wire axis is expected at the remanent state in cylindrical nanowires because of the strong longitudinal shape anisotropymuch stronger that the magnetocrystalline anisotropyof such 1D nanostructures that forces the magnetization to be oriented parallel to the longest dimension to minimize the magnetostatic energy. Similar magnetization configuration was observed in narrower cylindrical FeCoCu NWs in which the diameter modulation is 22 and 35 nm.30 On the other hand, the magnetic profiles of Figure 4i present a piecewise tendency with a linear dependence inside the NW and small variations outside of it. Inside the wire, both profiles exhibit a positive slope, while outside of the wire, a clear difference can be observed between the magnetic profiles in both segments: a positive variation occurs in the profile extracted from the outer part of the SD segment, while a negative variation is obtained around the LD segment. The magnetic phase shift observed outside the NW is caused by the magnetic stray field spreading out of it, and the different variation of the phase shift along the y-direction suggests a change of the local stray field orientation from a SD segment to a LD segment. For mapping the in-plane projected stray field, we reconstruct magnetic flux images using a higher amplifier factor of n = 8 (see Figure 4d,h). We notice that the DM geometry of the FeCoCu NWs induces a complex demagnetizing field configuration where the stray field emerges/enters 9673

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano

Figure 6. 3D representation of the simulated remanent magnetic state for the DM FeCoCu NW of Figure 8. Colors inside the NW indicate the orientation of the x- and y-component of the normalized magnetization (mx and my, respectively). Colored arrows surrounding the NW represent the stray field (Hdx and Hdy) for an intermediate layer of the simulation (at a height of z = 120 nm). (b−e) A 3D perspective view of the remanent spin configuration for some sections of the simulated DM NW: (b) cross-section around a DM transition, (c) intermediate section (black dashed box in (a)) that covers one SD-LD segment, (c) SD NW end (red dashed box in (a)), and (d) LD NW end (green dashed box in (a)). Colors in (b−e) represent the my magnitude. Green arrows help to see the chirality of each vortex-like state.

shown in Figure 6 (videos of these 3D representations, viewed at different angles, are available as Movie S1, S2, and S3). Figure 6a displays colormap representations of the x- and ycomponent of the normalized magnetization, mx and my, respectively (mx = Mx/MS; my = My/MS, where the magnetization M = (Mx, My, Mz)), for the spins located at the surface of the NW and viewed from the +z-axis. Similar to that observed by EH experiments, a single color tone (red) in the mx colormap reveals that the remanent state of the NW tends to form a pseudo single-domain state with the spins mainly oriented along the +x-direction, except in the extremes where a white color tone indicates that the averaged spin orientation is perpendicular to the NW axis. This fact is easily observed in the my colormap where an important contribution of this component is found in both NW ends, causing intense blue and red tones. In addition, the my colormap also reveals that the longitudinal alignment of the magnetization is also altered in the LD segments, where red and/or blue tones are also present. Making a cross-sectional representation of my around one of the diameter-variation crossovers (see Figure 6b), we found that these regions favor the formation of vortex-like structures, deviating the spin orientation up to 45° with respect to the xaxis at the surface of the NW where the tilting effect is higher. As we see in Figure 6c, this curling state is extended only inside the LD segments, and its effect is attenuated as it moves away

features. In the thinner NW end, the parallel and longitudinal alignment of the magnetic flux observed close to the diametervariation crossover (right-side of Figure 5b,d) is slightly distorted by a progressive increase of the line spacing as they approach the lateral edge. In the thicker NW end, the magnetic flux is strongly modified at the lateral edge of the end, forming a type of C-shaped flux. Outside of both ends, the stray field distribution resembles that of 1D nanostructures with a uniaxial magnetic domain.43 As a preliminary interpretation of the 2D projection magnetic induction inside the NW ends, we could conclude that the longitudinal spin alignment is only altered at the very end of the wires, being much stronger in thicker ends. In fact, the formation of micromagnetic closure structures (here labeled as C-shaped) is usual in NWs and appears to reduce the net perpendicular component of the magnetization at the magnetic/nonmagnetic border (i.e., to reduce the stray field energy in that border).35,44 Moreover, depending on the actual shape of the NW end, a diverse local distribution of magnetic lines is expected. The micromagnetic simulations permit us not only to validate the interpretation of the EH results but also to go deeper in the full description of this remanent state thanks to the 3D capability to reconstruct the spin configuration and the demagnetizing field. A summary of 3D representations of the resulting simulated remanent state for the NW of Figure 8 is 9674

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano

Figure 7. Colormap and arrows corresponding to the 2D demagnetizing field [Hdxy = (Hdx, Hdy)] induced by the remanent magnetic state of the NW in an intermediate plane (z = 120 nm). Color scale represents the magnitude of Hdx, and the green line delineates the NW contour. (b) Hdx and Hdy profiles taken along the red arrow.

between a negative charge (at the thin end) and positive charge (in the neighboring DM transition, to the right). Similar to the diameter-variation crossover, the ends of the wire also induce the formation of vortex-like structures in the lateral edges, and it could extend along the entire end (in the LD one). In the SD end, the vortex state is only formed in the edge, and it disappears as the spins are further away of it, aligning parallel to the NW axis. Compared with those observed in the diameter-variation crossover, the vortex structures at the ends of the wire induce a curling state with spins highly tilted toward the perpendicular direction of the wire (up to 90° respect to the x-axis for spins at the NW surface), and this type of flux-closure magnetization state at the NW ends favors the minimization of magnetic charges. The complexity of the magnetic charges in DM FeCoCu NWs also induces a nonintuitive behavior in the demagnetizing field inside the wire. The alternating ordering of positive and negative magnetic charges not only favors a parallel flux of the stray field to the magnetization around SD segments but also creates an internal demagnetizing field parallel to the magnetization. This fact can be visualized in Figure 7a where a 2D vector map of the demagnetizing field (Hd) [Hd(x, y, z) = (Hdx, Hdy, Hdz)] for an intermediate xy plane is displayed. A reddish tone associated with Hdx is observed inside the SD segments indicating that the internal dipolar field is parallel to the magnetization. The magnitude and variation of Hd at such a xy plane can be observed in the plot of Figure 7b. In an intermediate section of the wire away from the ends, we see as the internal dipolar field is mainly aligned along the NW axis, and its magnitude in the SD segments decreases dramatically as we move to the center of the segment, passing from 96 mT (diameter-variation crossover) to 27 mT (center of the SD segment); inside the LD segment, the dipolar field magnitude exhibits a sinusoidal-like behavior with a maximum value of 81 mT at the center of the segment and two minimum values of 49 mT in intermediate positions. This peculiar behavior could play an important role in the DW motion process through SD segments because a parallel alignment of the magnetization and

from the diameter-variation crossover. Similar to the magnetic charge sign, the chirality of these vortex structures changes consecutively along the wire: they are correlated as each positive (negative) magnetic charge contains a clockwise (counterclockwise) vortex-like chirality structure. This finding is in good agreement with those found in narrower DM FeCoCu NWs30 and also in bamboo-like FeCoCu NWs (DM NWs where the LD segment length is much shorter than the SD segment ones).23,45 The formation of vortex structures in the DM transition is induced by the system in an attempt to minimize the magnetic surface charge density incited by the diameter variation. However, the surface charge density is not reduced to zero, so the DM transitions turn into magnetic poles whereby the stray field comes in and out of the NW. This magnetic behavior enables the bright and dark contrast observed in the MFM image of Figure 3b. Moreover, the alternating bright and dark contrast pattern is a direct consequence of the way the diameter-variation crossovers act as alternating source and sink of magnetic fields even without the creation of any domain within the NW. As these vortex-like structures only extend along LD segments, they control their local magnetization configuration, creating a kind of torsional deformed longitudinal alignment of the magnetization because of the chirality difference of the vortex-like states located in their limits. The torsional deformed structure of the spins causes the bending effect of the 2D projection flux lines inside the LD segments. Contrary, the SD segments are not affected by the vortex states, so the spins are perfectly aligned along the NW axis, as we see in Figure 6c. A particular behavior is observed in the LD segment next to the thin end where the vortex state is extended along the entire segment because it is enclosed by two vortex structures with the same chirality. This peculiar effect is linked to a magnetic charge frustration of the DM transition located at the left-side extremity of this particular LD segment. The magnetic charge frustration occurs because this specific DM transition is 9675

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano

Figure 8. (Right) A perspective view of the 3D geometry used in the micromagnetic simulation. Gray region represents the magnetized body (NW) and the transparent box indicates the universe of the simulation. Black shadow corresponds to the xy plane projection of the NW. (Left) Zoom of a part of the black shadow to see the lateral dimension of the universe of the simulation at a half height (z = 120 nm).

the demagnetizing field minimizes even more the magnetostatic energy of the system and could induce a coherent magnetization switching where all the spins rotate coherently, being forced to be collinear between them during the switching. A coherent-mode rotation in the thinnest segment has been previously observed in DM NWs of Ni80Fe20.20

Beyond DW motion-based technologies, the artificial fabrication of local magnetic charges in DM NWs opens an alternative route for storing and writing information. While a precise control of the position and sign of each magnetic charge into a DM NW will allow storing data as a binary system (0s and 1s will be determined by the magnetic charge signs), the local magnetic fields created by the magnetic charges can be used to change the local magnetic states of 1D soft magnetic materials located very close to the DM NW, and thus we achieve printing the stored information in the magnetic charge array. Furthermore, the local field of the magnetic charge can be used as a trap for pinning DWs that are propagated in a nearby nanowire.46−49

CONCLUSIONS The nanoscale ability of EH to perform a quantitative analysis of the magnetic configuration in ferromagnetic nanostructures, supported by micromagnetic simulations, allowed a detailed characterization of the local spin and demagnetizing field configuration of cylindrical FeCoCu NWs with alternating modulation in the diameter. At remanence, the DM geometry of the wires studied here (diameter variation from 144 to 100 nm) alters the longitudinal alignment of the magnetization by inducing vortex-like states, as well as magnetic charges, in regions where the diameter is varied. These vortex-like structures play an important role in the spin configuration of LD segments because they provoke a torsionally deformed alignment of the spins in these thick segments. On the other hand, the induced magnetic charges control the local demagnetizing field of DM FeCoCu NWs where the magnetic interaction between neighboring charges and the alternating positive/negative charge arrangement along the wire promotes a flux-closure configuration around LD segments and a dipolar field that flows parallel to the magnetization both inside and around of SD segments. The complete description of the remanent magnetic configuration carried out by EH and micromagnetic simulation allowed clarifying the origin of bright and dark magnetic contrasts observed in MFM measurements that resemble to DWs. These magnetic contrasts are produced by the outgoing/incoming stray field of the magnetic charges. This finding suggests that future studies for the control of the DW motion in DM cylindrical NWs will be focused on how the vortex-like structures and magnetic charges affect the magnetization reversal process. Ideally, we expect that both magnetic phenomena, or one of them, act as a nucleation and pinning center of DWs. As they are induced by the DM geometry of wire, their optimization for handling DWs exclusively depends on the ability to grow nanowires with a specifically tailored modulation.

EXPERIMENTAL DETAILS Cylindrical DM FeCoCu NWs were prepared by filling self-assembled pores of anodic aluminum oxide (AAO) templates by electroplating. The templates were obtained by pulsed hard anodization in oxalic aqueous solution (0.3M) containing 5 vol % ethanol at a constant temperature of 0 °C. During the anodization, a constant voltage of 80 V was first applied for 600 s to produce a protective aluminum oxide layer at the surface of the disc, which avoids breaking or burning effects during subsequent pulsed hard anodization. After that, the voltage was slowly increased (0.05 V/s) up to 100 V and kept constant for 400 s, which ensures the alignment of the nanochannels. Nanopores with tailored periodical diameter modulations were produced by applying pulses of 130 and 100 V for 5 and 150 s, respectively. The pulses were repeated 30 times to obtain a total length of the modulated pores of about 50 μm. The resulting cylindrical pores are formed by segments with diameters of about 100 and 140 nm, respectively, while the center-to-center interpore distance is kept constant at 320 nm. Finally, the residual Al and the alumina barrier layer at the bottom of the foils were chemically etched, and an Au layer was sputtered to serve later as an electrode for final electroplating of nanowires.22 Fe28Co67Cu05 alloys (referred as FeCoCu in this work) were grown into the modulated pores of AAO templates, at room temperature, by DC electrodeposition using the electrolyte: 0.12 M CoSO4, 0.05 M FeSO4, 0.01 M CuSO4, 0.16 M H3BO3, and 0.06 M C6H8O6. The NWs were then released from the alumina membrane to proceed with the characterization of individual elements, deposing them onto a silicon substrate and a TEM carbon grid to perform MFM and TEM measurements, respectively. Studies of the morphology and magnetic distribution of DM FeCoCu NWs were performed by HRTEM and EH techniques, respectively, using a Hitachi HF-3300 (I2TEM-Toulouse) microscope 9676

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano operated at 300 kV. This dedicated microscope, specially optimized to perform interferometry experiments, is equipped with a cold field emission gun that emits highly coherent electrons with a very high brightness of about ∼109 A/cm2·sr50 and a spherical aberration corrector (aplanator B-COR from CEOS) to correct for the off-axial aberration in both TEM and Lorentz modes. EH experiments were carried in a corrected Lorentz mode using the normal stage of the I2TEM (the sample is placed within the pole piece of the objective lens, which is switched off to study the magnetic NWs in a field-free condition) that allows reaching a very large field of view up to 1.3 μm, with a fringe spacing of the holograms of 1.72 nm. In addition, a double-biprisms setup was used to avoid the Fresnel fringes created in the holograms when only one biprism is used. In addition, a comparative analysis between the EH results with MFM measurements was done. Topography and MFM images, at remanence, were obtained in a Cervantes system from Nanotec Electronica, using a twopass operation mode. Commercial MFM probes from Nanosensors were used in these experiments. Static micromagnetic simulations of the remanent state were performed using the OOMMF code51 and the following magnetic parameter: saturation magnetization (experimentally determined by EH), μ0MS = 1.33 T ≈ 1060 kA/m; exchange constant, A = 26× 10−12 J/m. As a good approximation, the anisotropy constant was neglected (K = 0) due to the polycrystalline character of the NWs. For these micromagnetic simulations, a representative 3D DM NW of 2.6 μm length was built by stacking a magnetic unit cell of 5 × 5 × 5 nm3. To simulate the stray fields surrounding the nanowire, lateral dimensions (along the x and y directions) larger than the diameter of the nanowire were chosen, leaving an empty space of about 100−200 nm between the NW surfaces and the universe frontiers. The geometrical parameters such as the size and shape of the DM NW as well as the size of the simulation universe are schematically illustrated in Figure 8. To perform the simulation calculations, the NW is set in an initial magnetization state where all spins are perfectly oriented along the xaxis; the static simulation then runs by using a second-order predictor corrector until the total energy of the system reaches a defined minimum.

measurements. L.A.R. and E.S. wrote the preliminary manuscript. Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENTS This work was supported by the European Union Seventh Framework Program under a contract for an Integrated Infrastructure Initiative reference no. 312483-ESTEEM2, by the French National Research Agency under the “Investissement d’Avenir” program reference no. ANR-10-EQPX-38-01, and by the project MAT2013-48054-C2-1-R from the MINECO-Spain. The authors acknowledge the “Conseil Regional Midi-Pyrénées” and the European FEDER for financial support within the CPER program. The authors also gratefully acknowledge Dr. C. Magén and Dr. P. Algarabel for facilitating the use of their high performance computing server to perform the micromagnetic simulations. REFERENCES (1) Allwood, D. A.; Xiong, G.; Faulkner, C. C.; Atkinson, D.; Petit, D.; Cowburn, R. P. Magnetic Domain-Wall Logic. Science 2005, 309, 1688−1692. (2) Allwood, D. A.; Xiong, G.; Cowburn, R. P. Writing and Erasing Data in Magnetic Domain Wall Logic Systems. J. Appl. Phys. 2006, 100, 123908. (3) Parkin, S. S. P.; Hayashi, M.; Thomas, L. Magnetic Domain-Wall Racetrack Memory. Science 2008, 320, 190−194. (4) Hayashi, M.; Thomas, L.; Moriya, R.; Rettner, C.; Parkin, S. S. P. Current-Controlled Magnetic Domain-Wall Nanowire Shift Register. Science 2008, 320, 209−211. (5) Corte-León, H.; Krzysteczko, P.; Schumacher, H. W.; Manzin, A.; Cox, D.; Antonov, V.; Kazakova, O. Magnetic Bead Detection Using Domain Wall-Based Nanosensor. J. Appl. Phys. 2015, 117, 17E313. (6) Corte-León, H.; Gribkov, B.; Krzysteczko, P.; Marchi, F.; Motte, J.-F.; Schumacher, H. W.; Antonov, V.; Kazakova, O. Magnetic Scanning Gate Microscopy of a Domain Wall Nanosensor Using Microparticle Probe. J. Magn. Magn. Mater. 2016, 400, 225−229. (7) Bran, C.; Palmero, E. M.; del Real, R. P.; Vazquez, M. CoFeCu Electroplated Nanowire Arrays: Role of Composition and Annealing on Structure and Magnetic Properties. Phys. Status Solidi A 2014, 211, 1076−1082. (8) Ivanov, Y. P.; Alfadhel, A.; Alnassar, M.; Perez, J. E.; Vazquez, M.; Chuvilin, A.; Kosel, J. Tunable Magnetic Nanowires for Biomedical and Harsh Environment Applications. Sci. Rep. 2016, 6, 24189. (9) Koyama, T.; Chiba, D.; Ueda, K.; Kondou, K.; Tanigawa, H.; Fukami, S.; Suzuki, T.; Ohshima, N.; Ishiwata, N.; Nakatani, Y.; Kobayashi, K.; Ono, T. Observation of the Intrinsic Pinning of a Magnetic Domain Wall in a Ferromagnetic Nanowire. Nat. Mater. 2011, 10, 194−197. (10) Piao, H.-G.; Shim, J.-H.; Djuhana, D.; Kim, D.-H. Intrinsic Pinning Behavior and Propagation Onset of Three-Dimensional Bloch-Point Domain Wall in a Cylindrical Ferromagnetic Nanowire. Appl. Phys. Lett. 2013, 102, 112405. (11) Junyeon, K.; Myung-Hwa, J.; Sug-Bong, C. Stability of Magnetic Domains With Notches in Permalloy Nanowires. IEEE Trans. Magn. 2009, 45, 2481−2484. (12) Noh, S. J.; Miyamoto, Y.; Okuda, M.; Hayashi, N.; Keun Kim, Y. Effects of Notch Shape on the Magnetic Domain Wall Motion in Nanowires with in-Plane or Perpendicular Magnetic Anisotropy. J. Appl. Phys. 2012, 111, 07D123. (13) Rodríguez, L. A.; Magén, C.; Snoeck, E.; Gatel, C.; Marín, L.; Serrano-Ramón, L.; Prieto, J. L.; Muñoz, M.; Algarabel, P. A.; Morellon, L.; De Teresa, J. M.; Ibarra, M. R. Quantitative in Situ Magnetization Reversal Studies in Lorentz Microscopy and Electron Holography. Ultramicroscopy 2013, 134, 144−154.

ASSOCIATED CONTENT S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsnano.6b05496. Basic description of electron holography; magnetic reconstruction for a long section of a wire; bending effect of the magnetic flux lines in the LD segments (PDF) Rotate animation of the simulated magnetization distribution of Figure 6a using as color representation the x-component of the magnetization (AVI) Rotate animation of the simulated magnetization distribution of Figure 6a using as color representation the y-component of the magnetization (AVI) Rotate animation of the simulated magnetization distribution of Figure 6c using as color representation the y-component of the magnetization (AVI)

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Author Contributions

L.A.R. and D.R. performed the EH experiments. L.A.R. and C.G. performed the data treatment of the EH experiments. M.V., C.B., and A.A. contributed to the conception and design of the NWs. C.B. fabricated the NWs. L.A.R. performed the micromagnetic simulations. E.B and A.A. performed the MFM 9677

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678

Article

ACS Nano (14) Brandão, J.; Novak, R. L.; Lozano, H.; Soledade, P. R.; Mello, A.; Garcia, F.; Sampaio, L. C. Control of the Magnetic Vortex Chirality in Permalloy Nanowires with Asymmetric Notches. J. Appl. Phys. 2014, 116, 193902. (15) Yuan, H. Y.; Wang, X. R. Domain Wall Pinning in Notched Nanowires. Phys. Rev. B: Condens. Matter Mater. Phys. 2014, 89, 054423. (16) Chen, T.-C.; Kuo, C.-Y.; Mishra, A. K.; Das, B.; Wu, J.-C. Magnetic Domain Wall Motion in Notch Patterned Permalloy Nanowire Devices. Phys. B 2015, 476, 161−166. (17) Yuan, H. Y.; Wang, X. R. Boosting Domain Wall Propagation by Notches. Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 92, 054419. (18) Allende, S.; Altbir, D.; Nielsch, K. Magnetic Cylindrical Nanowires with Single Modulated Diameter. Phys. Rev. B: Condens. Matter Mater. Phys. 2009, 80, 174402. (19) Salem, M. S.; Sergelius, P.; Corona, R. M.; Escrig, J.; Görlitz, D.; Nielsch, K. Magnetic Properties of Cylindrical Diameter Modulated Ni80Fe20 Nanowires: Interaction and Coercive Fields. Nanoscale 2013, 5, 3941. (20) Tejo, F.; Vidal-Silva, N.; Espejo, A. P.; Escrig, J. Angular Dependence of the Magnetic Properties of Cylindrical Diameter Modulated Ni80Fe20 Nanowires. J. Appl. Phys. 2014, 115, 17D136. (21) Wong, D. W.; Chandra Sekhar, M.; Gan, W. L.; Purnama, I.; Lew, W. S. Dynamics of Three-Dimensional Helical Domain Wall in Cylindrical NiFe Nanowires. J. Appl. Phys. 2015, 117, 17A747. (22) Pitzschel, K.; Bachmann, J.; Martens, S.; Montero-Moreno, J. M.; Kimling, J.; Meier, G.; Escrig, J.; Nielsch, K.; Görlitz, D. Magnetic Reversal of Cylindrical Nickel Nanowires with Modulated Diameters. J. Appl. Phys. 2011, 109, 033907. (23) Bran, C.; Berganza, E.; Palmero, E. M.; Fernandez-Roldan, J. A.; Del Real, R. P.; Aballe, L.; Foerster, M.; Asenjo, A.; Fraile Rodríguez, A.; Vazquez, M. Spin Configuration of Cylindrical Bamboo-like Magnetic Nanowires. J. Mater. Chem. C 2016, 4, 978−984. (24) Palmero, E. M.; Bran, C.; del Real, R. P.; Vázquez, M. Vortex Domain Wall Propagation in Periodically Modulated Diameter FeCoCu Nanowire as Determined by the Magneto-Optical Kerr Effect. Nanotechnology 2015, 26, 461001. (25) Lee, W.; Ji, R.; Gösele, U.; Nielsch, K. Fast Fabrication of LongRange Ordered Porous Alumina Membranes by Hard Anodization. Nat. Mater. 2006, 5, 741−747. (26) Sulka, G. D.; Brzózka, A.; Liu, L. Fabrication of DiameterModulated and Ultrathin Porous Nanowires in Anodic Aluminum Oxide Templates. Electrochim. Acta 2011, 56, 4972−4979. (27) Esmaeily, A. S.; Venkatesan, M.; Razavian, A. S.; Coey, J. M. D. Diameter-Modulated Ferromagnetic CoFe Nanowires. J. Appl. Phys. 2013, 113, 17A327. (28) Prida, V. M.; García, J.; Hernando, B.; Bran, C.; Vivas, L. G.; Vázquez, M. Electrochemical Synthesis of Magnetic Nanowires with Controlled Geometry and Magnetic Anisotropy. In Magnetic Nanoand Microwires; Design, Synthesis, Properties and Applications; Vázquez, M., Eds; Woodhead Publishing: Cambridge, U.K., 2015; pp 41−104. (29) Yan, M.; Kákay, A.; Gliga, S.; Hertel, R. Beating the Walker Limit with Massless Domain Walls in Cylindrical Nanowires. Phys. Rev. Lett. 2010, 104, 057201. (30) Iglesias-Freire, Ó .; Bran, C.; Berganza, E.; Mínguez-Bacho, I.; Magén, C.; Vázquez, M.; Asenjo, A. Spin Configuration in Isolated FeCoCu Nanowires Modulated in Diameter. Nanotechnology 2015, 26, 395702. (31) Biziere, N.; Gatel, C.; Lassalle-Balier, R.; Clochard, M. C.; Wegrowe, J. E.; Snoeck, E. Imaging the Fine Structure of a Magnetic Domain Wall in a Ni Nanocylinder. Nano Lett. 2013, 13, 2053−2057. (32) Reyes, D.; Biziere, N.; Warot-Fonrose, B.; Wade, T.; Gatel, C. Magnetic Configurations in Co/Cu Multilayered Nanowires: Evidence of Structural and Magnetic Interplay. Nano Lett. 2016, 16, 1230−1236. (33) Midgley, P. A.; Dunin-Borkowski, R. E. Electron Tomography and Holography in Materials Science. Nat. Mater. 2009, 8, 271−280. (34) Wolf, D.; Lubk, A.; Röder, F.; Lichte, H. Electron Holographic Tomography. Curr. Opin. Solid State Mater. Sci. 2013, 17, 126−134.

(35) Wolf, D.; Rodriguez, L. A.; Béché, A.; Javon, E.; Serrano, L.; Magen, C.; Gatel, C.; Lubk, A.; Lichte, H.; Bals, S.; Van Tendeloo, G.; Fernández-Pacheco, A.; De Teresa, J. M.; Snoeck, E. 3D Magnetic Induction Maps of Nanoscale Materials Revealed by Electron Holographic Tomography. Chem. Mater. 2015, 27, 6771−6778. (36) Minguez-Bacho, I.; Rodriguez-López, S.; Vázquez, M.; Hernández-Vélez, M.; Nielsch, K. Electrochemical Synthesis and Magnetic Characterization of Periodically Modulated Co Nanowires. Nanotechnology 2014, 25, 145301. (37) Palmero, E. M.; Bran, C.; del Real, R. P.; Magen, C.; Vazquez, M. Structural and Magnetic Characterization of FeCoCu/Cu Multilayer Nanowire Arrays. IEEE Magn. Lett. 2014, 5, 1−4. (38) Vock, S.; Hengst, C.; Wolf, M.; Tschulik, K.; Uhlemann, M.; Sasvári, Z.; Makarov, D.; Schmidt, O. G.; Schultz, L.; Neu, V. Magnetic Vortex Observation in FeCo Nanowires by Quantitative Magnetic Force Microscopy. Appl. Phys. Lett. 2014, 105, 172409. (39) Viñas, S. L.; Salikhov, R.; Bran, C.; Palmero, E. M.; Vazquez, M.; Arvan, B.; Yao, X.; Toson, P.; Fidler, J.; Spasova, M.; Wiedwald, U.; Farle, M. Magnetic Hardening of Fe30Co70 Nanowires. Nanotechnology 2015, 26, 415704. (40) Bran, C.; Ivanov, Y. P.; García, J.; del Real, R. P.; Prida, V. M.; Chubykalo-Fesenko, O.; Vazquez, M. Tuning the Magnetization Reversal Process of FeCoCu Nanowire Arrays by Thermal Annealing. J. Appl. Phys. 2013, 114, 043908. (41) Lopatin, S.; Ivanov, Y. P.; Kosel, J.; Chuvilin, A. Multiscale Differential Phase Contrast Analysis with a Unitary Detector. Ultramicroscopy 2016, 162, 74−81. (42) Ivanov, Y. P.; Chuvilin, A.; Lopatin, S.; Kosel, J. Modulated Magnetic Nanowires for Controlling Domain Wall Motion: Toward 3D Magnetic Memories. ACS Nano 2016, 10, 5326−5332. (43) Engel-Herbert, R.; Hesjedal, T. Calculation of the Magnetic Stray Field of a Uniaxial Magnetic Domain. J. Appl. Phys. 2005, 97, 074504. (44) Dmytriiev, O.; Al-Jarah, U. a. S.; Gangmei, P.; Kruglyak, V. V.; Hicken, R. J.; Mahato, B. K.; Rana, B.; Agrawal, M.; Barman, A.; Mátéfi-Tempfli, M.; Piraux, L.; Mátéfi-Tempfli, S. Static and Dynamic Magnetic Properties of Densely Packed Magnetic Nanowire Arrays. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 87, 174429. (45) Berganza, E.; Bran, C.; Jaafar, M.; Vázquez, M.; Asenjo, A. Domain Wall Pinning in FeCoCu Bamboo-like Nanowires. Sci. Rep. 2016, 6, 29702. (46) Ahn, S.-M.; Moon, K.-W.; Cho, C.-G.; Choe, S.-B. Control of Domain Wall Pinning in Ferromagnetic Nanowires by Magnetic Stray Fields. Nanotechnology 2011, 22, 085201. (47) Mironov, V. L.; Ermolaeva, O. L.; Skorohodov, E. V.; Klimov, A. Y. Field-Controlled Domain Wall Pinning-Depinning Effects in a Ferromagnetic Nanowire-Nanoislands System. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 85, 144418. (48) Hiramatsu, R.; Koyama, T.; Hata, H.; Ono, T.; Chiba, D.; Fukami, S.; Ishiwata, N. Domain Wall Pinning by a Stray Field from NiFe on a Co/Ni Nanowire. J. Korean Phys. Soc. 2013, 63, 608−611. (49) Mironov, V. L.; Ermolaeva, O. L. Domain Wall Pinning in a Ferromagnetic Nanowire by Stray Fields of Nanoparticles. Bull. Russ. Acad. Sci.: Phys. 2014, 78, 16−20. (50) Houdellier, F.; de Knoop, L.; Gatel, C.; Masseboeuf, A.; Mamishin, S.; Taniguchi, Y.; Delmas, M.; Monthioux, M.; Hÿtch, M. J.; Snoeck, E. Development of TEM and SEM High Brightness Electron Guns Using Cold-Field Emission from a Carbon Nanotip. Ultramicroscopy 2015, 151, 107−115. (51) Donahue, M. J.; Porter, D. G. OOMMF User’s Guide, version 1.0; Interagency Report NISTIR 6376; National Institute of Standards and Technology: Gaithersburg, MD, 1999.

9678

DOI: 10.1021/acsnano.6b05496 ACS Nano 2016, 10, 9669−9678