Mapping the domain wall pinning landscape in ferromagnetic films

arXiv:1510.07059v1 [cond-mat.mes-hall] 23 Oct 2015

Robert Badea and Jesse Berezovsky∗ Department of Physics, Case Western Reserve University, Cleveland, OH E-mail: [email protected]

Abstract

lating about a central core 7–9 . The vortex core has a half-width ∼ 10 nm, over which range the magnetization is oriented largely out of plane 10 . The vortex core can be seen as a zero-dimensional analog of a one-dimensional domain wall, essentially acting like a scanning probe for imaging the pinning potential 7,11 . The interaction between a vortex domain and a pinning site has been characterized for both strong, artificially created pinning sites 12–16 , and weaker naturally-occurring pinning sites 11,17–20 . It was demonstrated that single point defects were capable of trapping the vortex core and thus suppressing its motion in response to applied magnetic field. While some information about the pinning potential can be extracted from these measurements, the techniques for data collection and analysis have generally been too cumbersome to extract the type of pinning potential maps presented here. For small displacements, the motion of the vortex core in an in-plane magnetic field H can be described by the rigid vortex model 21 , via a potential U0 (r, H) ≈ 21 k|r|2 −kχ0 (Hy x+Hx y), where r is the in-plane displacement of the core from the center and k is the stiffness of the vortex domain which depends on the demagnetization factor of the disk. This potential describes a paraboloid with the position of the minimum shifted perpendicular to an applied in-plane magnetic field. Vortex pinning can be included in the model by adding localized potential minima U p,i (r) 17,22 . As a magnetic field Hx is increased, the vortex core moves roughly in a line in the y-direction, with deviations from the line caused by the particular

The propagation of domain walls in a ferromagnetic film is largely determined by domain wall pinning at defects in the material. In this letter we map the effective potential landscape for domain wall pinning in Permalloy films by raster scanning a single ferromagnetic vortex domain and monitoring the hysteretic vortex displacement vs. applied magnetic field. The measurement is carried out using a differential magneto-optical microscopy technique which yields spatial sensitivity ∼ 10 nm. We present a simple algorithm for extracting an effective pinning potential from the measurement of vortex displacement vs. applied field. The resulting maps of the pinning potential reveal distinct types of pinning sites, which we attribute to zero-, one-, and two-dimensional defects in the Permalloy film. Proposed spintronic storage devices 1,2 and logic elements 3 rely on the controlled movement of domain walls in micromagnetic structures. This motion is strongly affected by pinning of the domain walls at defects 4–6 . Realization of this technology requires a deeper understanding of the microscopic origins of pinning and tools to characterize the pinning landscape. A ferromagnetic (FM) vortex provides a nanoscale probe for mapping the domain wall pinning properties of a film. In a thin, micron-scale FM disk lacking magnetocrystalline anisotropy, the ground state of the magnetization is a single vortex domain with in-plane magnetization circu∗ To

whom correspondence should be addressed

1

arrangement of pinning sites. It is convenient to represent the potential along this vortex path as an effective 1D potential u(y, Hx ) = u0 (y, Hx ) + ∑ u p,i (y), with u0 = 12 ky2 − kχ0 Hx y. For example, Fig. 1d shows a possible effective 1D potential u vs. y cycling between two values of Hx , with two Gaussian pinning sites. At temperature T = 0, the vortex core resides at y0 , a local minimum of u, translating along with that minimum as Hx is changed. The rate of displacement with applied field χ = dy0 /dHx depends inversely on the curvature u00 at the local minimum. By definition, χ is suppressed when the vortex is pinned as compared to the unpinned case where χ = χ0 . Sudden changes in the pinning configuration occur with a change in Hx when the presently occupied local minimum disappears (as illustrated in Fig. 1d). Assuming the pinning sites are sufficiently spaced to be treated independently, the vortex leaves the ith pinning site when the maximum or minimum slope of u p,i is cancelled by the slope of u0 , which occurs at pairs of values (y± , Hx± ) that satisfy u0 (y± , Hx± ) = u00p,i (y± , Hx± ) = 0.

(iv)

b

Permalloy Disk

d 2 1 0 -1

Microscope Objective

(iii)

c ∆y0 (arb.)

(v) (vi)

Ha Hb

Hx (kA/m)

a 3 2 1 0 -1 -2 -3 (i) (ii)

θ

H=Ha

u(y)

y

H=Hb

-2 Time

Figure 1: Measuring hysteretic vortex displacement. (a) Sweep of AC magnetic field Hx used to measure vortex displacement, with upper end (Ha ) and lower end (Hb ) of the 15-kHz square wave shown in blue and red respectively. Total measurement time ∼ 3 min. (b) Vortex displacement ∆y0 in response to Hx shown in (a). (c) Diagram of the experimental geometry (not to scale). (d) Illustration of an effective 1D potential u(y) with two pinning sites cycling between two fields Ha and Hb . The red dot indicates the vortex position.

croscopy apparatus 24 , described in more detail in Ref. 25 . The measurement is carried out in the longitudinal MOKE geometry via suitable masking of the probe laser which is focused through a high numerical aperture objective, and is therefore mainly sensitive to the magnetization component Mx . With the probe spot focused on the vortex core position, Mx = 0 by symmetry. As an in-plane field Hx is applied, a nonzero Mx appears which is roughly proportional to the vortex core displacement ∆y0 for small displacement. In order to observe the hysteretic features in the vortex motion arising from pinning, it is necessary to 1. achieve sufficient signal-to-noise to resolve changes in vortex position ∼ 10 nm and 2. carry out the measurement on sufficiently fast timescales to minimize thermal activation out of pinning sites. Both of these are accomplished using the differential nature of the experiment. A 15-kHz square wave current is driven through the stripline beneath the magnetic disks, producing a magnetic field Hx alternating between two values Ha and Hb . Using a lock-in technique, the measurement produces a signal θ ∝ Mx (Ha ) − Mx (Hb ) ∝ y0 (Ha ) − y0 (Hb ). The constant of proportionality between θ and vortex displacement is determined by measuring the vortex displacement in a static applied field by scanning the probe across the vortex 25 . Though

(1)

The ± indicates the two solutions with < 0 and u0p,i (y− , Hx− ) > 0. As Hx is swept through such a point, a sudden jump in y0 and/or χ is observed. As illustrated in Fig. 1d, multiple local minima may be present at a particular value of Hx , resulting in hysteresis in y0 as Hx is swept up and down. At T > 0, thermal activation of the vortex out of a local minimum is possible 23 , causing smearing of the sudden jumps and a shift in the critical field at which the jumps occur, reducing the observed hysteresis 11 . The samples measured here were Permalloy (Py, Ni0.81 Fe0.19 ) disks with diameter d = 2 µm and thicknesses t = 30 to 40 nm. Data from two samples, D1 and D2, are presented. Samples were were fabricated via electron-beam lithography, electron-beam evaporation, and liftoff atop a gold microstrip transmission line. Naturally occurring defects in the Py serve as pinning sites for the vortex core. The experimental setup, with geometry shown in Fig. 1c, is based on a standard, room temperature magneto-optical Kerr effect (MOKE) miu0p,i (y+ , Hx+ )

2

∆y0 (arb.)

a D1 3 Hy= 4.5 kA/m 2 1 0 -1 -2 -3

b D1 Hy= 0.8 kA/m

field decreasing to Hb . Thus segments (i) and (ii) yield the initial increasing and decreasing paths of a traditional hysteresis loop. The remaining segments reveal the rest of the hysteresis loop in a similar manner. By plotting vortex displacement ∆y0 relative to initial position vs. Hs , hysteresis loops are formed, as shown in Fig. 2. Figure 2a-c shows three characteristic ∆y0 vs. Hs loops. The data in a and b are from the same disk (D1) with the vortex shifted to different positions in the x-direction by changing Hy . The data in c is from a different disk (D2) on a completely separate sample. All three show multiple jumps in both ∆y0 and the slope χ as Hs is swept, corresponding to changes in the pinning state. When the vortex is not pinned, then χ = χ0 which is the maximum value of χ (except for at sudden jumps of ∆y0 ). Such a free region can be seen on the left side of Fig. 2b. When the vortex is pinned, χ is reduced, as is the case at Hs = 0 in both Fig. 2a and b. The value of χ indicates the curvature of the pinning potential with a larger χ corresponding to less curvature. From the type of data shown in Fig. 2a-c, we can construct an effective 1D potential that gives rise to the observed loops. As described above, jumps ± in y0 or χ occur at a set of points (y± i , Hx,i ) that satisfy Eq. 1. These points can be directly read off from the y0 vs. Hx data, and through Eq. 1 specify both the value of u0p,i and that u0p,i is a local maximum or minimum at that value of y. There may be additional solutions to Eq. 1 that do not appear in the data either because that value of y is jumped over, or the value of Hx is beyond the range of the scan. If there were a pinning site i where neither − − + solution (y+ i , Hx,i ) nor (yi , Hx,i ) contributed to the observed pinning, then this pinning site would not be included in the reconstructed effective potential. If only one of the two solutions appears in the data, then knowledge of the other solution is required to reconstruct the potential, and this is approximated within constraints described in the supporting material. These pairs of points specify the value and position of the maximum and minimum u0p,i for each pinning potential u p,i . The total effective 1D potential is then constructed by connecting these points using a quadratic spline. The details of the algorithm, and additional assumptions which must be made are described in the sup-

c D2 Hy= -2.6 kA/m

i ii iii iv v vi

-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 Hs (kA/m) Hs (kA/m) Hs (kA/m) d e f

up (eV)

0 -10 -20 -30

∆y0 (nm)

g

-250 0 250 y (nm)

h

-250 0 250 y (nm)

i

-250 0 250 y (nm)

200 0 -200 -400 -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 Hs (kA/m) Hs (kA/m) Hs (kA/m)

Figure 2: Extracting 1D pinning potentials from hysteresis loops. (a-c) Vortex displacement ∆y0 vs. Hs on samples D1 and D2 at different static Hy , following the sweep path shown in Fig. 1a. (d-f) Effective 1D pinning potentials u p (y) extracted from the data shown in (a-c). (g-i) Simulated vortex displacement ∆y0 vs. Hs using the potentials u p in (d-f).

each data point is collected over a time ∼ 1 s, the dwell time of the vortex at Ha or Hb is only 33 µs. Due to the exponential nature of thermal activation, this reduction in dwell time by more than 4 orders of magnitude roughly reduces the thermally surmountable barrier by about 10kB T . We measure the hysteretic vortex displacement ∆y0 vs. Hx by sweeping the amplitude and offset of the applied square-wave magnetic field as illustrated in Fig. 1a, with the blue and red lines indicating the upper level (Ha ) and lower level (Hb ) respectively. In each segment (i)-(vi), there is one swept end Hs of the alternating magnetic field, either Hs = Ha or Hs = Hb . The resulting measured signal θ ∝ ∆y0 is shown in Fig. 1b as the difference between the red and blue lines. The red and blue lines show the net vortex displacement from the initial position at Ha and Hb . In segment (i), θ represents the vortex displacement starting from the initial position with field increasing to Ha . In segment (ii), θ represents the vortex displacement starting from the endpoint of segment (i), with 3

c ★★



2





3 Hb (kA/m)

b

1

W

3

2 1

L

0

-4 -2 0 2 4 Hy (kA/m)

-500

W

P P

L

2 3

d

50 eV

1 ∆y0 (arb.)

Ha (kA/m)

a

0

500 -500 y (nm)

0

500

Figure 3: 2D maps of the pinning potential. (a) and (b) Path (i) and (ii) of vortex displacement ∆y0 vs. Ha and Hb , as static field Hy shifts the free vortex position in the x-direction. Examples of regions of Hy where the vortex is trapped by the same pinning site are indicated with †, separated by transitions labelled ?. (c) and (d) Maps of pinning potential from samples D1 and D2 produced from series of 1D effective potentials with static Hy = −5570 to 5570 A/m in 100 steps. Examples of three types of pinning sites are labelled P,W, and L.

Fig. 3a and b reveal hysteresis in ∆y0 . These maps show features (e.g. labelled with †) with qualitatively similar behavior over a range of Hy , which end in sudden transitions (e.g. labelled with ?). A particular † feature is associated with a particular pinning potential U p,i . There is a range of Hy over which the vortex is captured by U p,i as Hx is swept. Over this range, the 2D path of the vortex through U p,i shifts, resulting in a continuous change in the corresponding 1D effective poten± change tial u p,i . As a result, the critical fields Hx,i over that range, giving rise to the curvature of the † features. At some value of Hy , the path of the vortex does not approach U p,i sufficiently to capture the vortex, and instead the vortex is captured by a different pinning site, or remains free at the minimum of U0 . At these critical values of Hy , a sudden change in the ∆y0 scans appears, such as at the ? features. An image reflecting the 2D pinning potential can be generated by constructing the 1D effective potential u p at each value of Hy and plotting them with an offset, as shown in Fig. 3c and d for disks D1 and D2. The series of scans cover Hy = −5570 to 5570 A/m in 100 steps. Plotted in this way, the series of scans roughly shows the shape of the potential in the x-direction as well as the ydirection. More accurately, the evolution along the series shows how the path of the vortex changes to pass through different pinning sites as the minimum of U0 is shifted in the x-direction. Each step of Hy shifts the minimum of U0 by ∆x0 ≈ 12 nm.

porting material. The effective 1D potentials u p = ∑ u p,i constructed from the data in Fig. 2a-c are shown in Fig. 2d-f. Note that the quantities specified by the measured data are values and positions of maximum or minimum u0p,i . Since, for simplicity, we have used a quadratic spline to connect these known points, the particular shape of the pinning potential, including the depth of the wells, should be understood to be qualitative. Nevertheless, the constructed potential does contain the necessary detail to reproduce the observed data. To verify this, we use a gradient descent method on the reconstructed 1D effective potential to numerically simulate the vortex displacement vs. Hx as in the experiment. The resulting simulated hysteresis loops are shown in Fig. 2g-i. In each case, the salient features of the data are reproduced. The simulation does not take thermal effects into account, so does not capture some smearing of the jumps or reduction of hysteresis for small pinning sites. We map the pinning potential in 2D by collecting a series of hysteresis loops of ∆y0 as above, incrementing the unpinned equilibrium position x0 in the x-direction, using a static field Hy . Figure 3a and b show path (i) and (ii) of the hysteresis loops vs. Hx over the range Hy = −5570 to 5570 A/m, which corresponds to a translation ∆x0 ≈ 1.2 µm. As in Fig. 2, jumps in ∆y0 with increasing Hx yield ± the critical fields Hx,i as the vortex jumps into or out of the ith pinning site. Differences between 4

The maps show features that persist across a range ∆Hy , or a corresponding range ∆x0 . These features correspond to pinning sites U p,i , with ∆x0 limited by the disappearance of the local minimum or by the vortex path not approaching sufficiently close to be captured. Three types of pinning sites can be identified in Fig. 3c and d, with examples labelled P, W, and L. P features are characterized by half-widths γ ∼ 10 nm, depths δ ∼ 1 eV, and ∆x0 ∼ 100 nm. These are consistent with point-like defects in the Py, where γ is set by the vortex core half-width, and ∆x0 is set by the full range over which the vortex core has significant out-of-plane magnetization 10 . Similar pinning potentials have been previously inferred using torsional microresonators 11 , or by mapping the vortex gyrotropic frequency 17–20 . W features are wider and deeper than P features, with γ ∼ 50 nm, δ ∼ 20 eV, and ∆x0 ∼ 200 nm. Because γ is greater than the halfwidth of the vortex core, the defects giving rise to W features are likely to have non-negligible spatial extent ∼ 50 nm in both x− and y-directions. Pinning behavior similar to W features has been previously observed in artificially created pinning sites 12–15 . In contrast to the symmetric nature of both W and P features, L features appear to arise from line-like defects. These features have γ ∼ 10 nm, δ ∼ 10 eV, and very large ∆x0 ∼ 500 nm or more. Here, γ is consistent with a point-like defect in the y-direction, but the large value of ∆x0 indicates significant spatial extent in the x direction. Note that this measurement is only sensitive to L features oriented mainly along the x-axis: for a line defect aligned close to the yaxis, the vortex would be free to move along the potential minimum over the entire hysteresis scan, and would look essentially like the unpinned case. Line defects at intermediate angles would likely give rise to jumps in the ∆y0 scan as the vortex goes into or out of the pinning potential separated by free motion as it translates along the line. Such behavior may not be distinguishable from that of a pair of P features. L features are not likely to arise from grain boundaries, as the scale of ∆x0 exceeds the typical grain size in these samples (∼ 50 nm). Instead, they may arise from more significant linear defects, such as cracking or buckling due to thermal contraction after deposition. These results provide a connection between the

microscopic defects in a FM film, and how those defects cause domain wall pinning. This technique may be used in future work to understand how the fabrication process or post-fabrication treatment can be used to control the degree and type of pinning present in a sample.

Supporting material Analysis of the data and the technique used to extract pinning potentials. Acknowledgement This work was supported by DOE, award No. DE-SC008148.

Competing financial interests The authors declare no competing financial interests.

References (1) S. Parkin, M. Hayashi, and L.Thomas. Magnetic domain-wall racetrack memory. Science, 320:190, 2008. 10.1126/science.1145799. (2) S. Barnes, J. Ieda, and S. Maekawa. Magnetic memory and current amplification devices using moving domain walls. Applied Physics Letters 89: 122507, 2006. 10.1063/1.2354036. (3) D. Allwood, G. Xiong, C. Faulkner, D. Atkinson, D. Petit, and R. Crowburn. Magnetic domain-wall logic Science 309:1688– 1692, 2005. 10.1126/science.1108813. (4) H. Barkhausen. Zwei mit hilfe der neuen verstärker entdeckte erscheinungen. Physik Zeitschrift, 20:401–403, 1916. (5) Alex Hubert and Rudolf Schäfer. Magnetic Domains: The Analysis of Magnetic Microstructures. Springer-Verlag Berlin Heidelberg, Berlin, 1996. 10.1007/978-3-54085054-0.

5

(6) Gianfranco Durin and Stefano Zapperi. Chapter 3 - The barkhausen effect. In Giorgio Bertotti, Isaak D. Mayergoyz, editor, The Science of Hysteresis, pages 181 – 267. Academic Press, Oxford, 2006.

(14) M. Rahm, R. Höllinger, V. Umansky, and D.Weiss. Influence of point defects on magnetic vortex structures. Journal of Applied Physics, 95:6708, 2004. 10.1063/1.1667448. (15) T. Uhlig, M. Rahm, C. Dietrich, R. Hollinger, M. Heumann, D. Weiss, and J. Zweck. Shifting and pinning of a magnetic vortex core in a parmalloy dot by a magnetic field. Physical Review Letters, 95:237205, 2005. 10.1103/PhysRevLett.95.237205.

(7) A. Wachowiak, J. Wiebe, M. Bode, O. Pietzsch, M. Morgenstern, and R. Wiesendanger. Direct observation of internal spin structure of magnetic vortex cores. Science, 298:577– 580, 2002. 10.1126/science.1075302. (8) R.P. Cowburn, D.K. Koltsov, A.O. Adeyeye, and M.E. Welland. Single-domain circular nanomagnets. Physical Review Letters, 83:1042–1045, 1999. 10.1103/PhysRevLett.83.1042.

(16) K. Kuepper, L. Bishoff, Ch. Akhmadaliev, J. Fassbender, H. Stoll, K. Chou, A. Puzic, K. Fauth, D. Dolgos, G. Schültz, V. Waeyenberge, T. Tyliszczak, I. Neudecker, G. Woltersdorf, C. Back. Vortex dynamics in permalloy disks with artificial defects: Suppression of the gyrotropic mode Physical Review Letters, 90:062506, 2007. 10.1063/1.2437710.

(9) T. Shinjo, T. Okuno, R. Hassdorf, K. Shigeto, and T. Ono. Magnetic vortex core observation in circular dots of permalloy. Science, 289:930–932, 2000. 10.1126/science.289.5481.930.

(17) T.Y. Chen, A.T. Galkiewicz, and P.A. Crowell. Phase diagram of magnetic vortex dynamics. Physical Review B, 85:180406, 2012. 10.1103/PhysRevB.85.180406.

(10) R. Höllinger, A. Killinger, and U. Krey. Statics and fast dynamics of nanomagnets with vortex structure. Journal of Magnetism and Magnetic Materials, 261:178– 189, 2003. 10.1016/S0304-8853(02)014713.

(18) T.Y. Chen, M.J. Erickson, P.A. Crowell, and C.Leighton. Surface roughness dominated pinning mechanisms of magnetic vortices in soft ferromagnetic films. Physical Review Letters, 109:097202, 2012. 10.1103/PhysRevLett.109.097202.

(11) J.A.J. Burgess, A.E. Fraser, F. Fani Sani, D. Vick, B.D. Hauer, J.P. Davis, and M.R. Freeman. Quantitative magnetomechanical detection and control of the barkhausen effect. Science, 339:1051–4, 2013. 10.1126/science.1231390.

(19) R.L. Compton and P.A. Crowell. Dynamics of a pinned magnetic vortex. Physical Review Letters, 97:137202, 2006. 10.1103/PhysRevLett.97.137202.

(12) M. Rahm, J. Biberger, V. Umansky, and D. Weiss. Vortex pinning at individual defects in magnetic nanodisks. Journal of Applied Sciences, 93:7429–7431, 2003. 10.1063/1.1558255.

(20) R.L. Compton, T.Y. Chen, and P.A. Crowell. Magnetic vortex dynamics in the presence of pinning. Physical Review B, 81:144412, 2010. 10.1103/PhysRevB.81.144412.

(13) M. Rahm, J. Stahl, W. Wegscheider, and D. Weiss. Multistable switching due to magnetic vortices pinned at artificial pinning sites. Applied Physics Letters, 85:1553, 2004. 10.1063/1.1785281.

(21) K. Guslienko, N. Novosad, Y. Otani, H. Shima, and F. Kukamichi. Magnetization reversal due to vortex nucleation, displacement, and annihilation in submicron ferromagnetic dot arrays. Physical Review B, 65:024414, 2001. 10.1103/PhysRevB.65.024414. 6

(22) J.A.J. Burgess, J.E. Losby, and M.R. Freeman. An analytical model for vortex core pinning in a micromagnetic disk. Journal of Magnetism and Magnetic Materials, 361:140–149,2014. 10.1016/j.jmmm.2014.02.078. (23) M. Eltschka, M. Wötzel, J. Rhensius, S. Krzyk, U. Nowak, M. Kläui, T. Kasama, R.E. Dunin-Borkowski, L.J. Heyderman, H.J. van Driel, and R.A. Duine. Nonadiabatic spin torque investigated using thermally activated magnetic domain wall dynamics. Physical Review Letters, 105:056601, 2010. 10.1103/PhysRevLett.105.056601. (24) M.R. Freeman and W.K. Hiebert. Stroboscopic microscopy of magnetic dynamics. In B. Hillebrands and K. Ounadjela, editors, Spin Dynamics in Confined Magnetic Structures I. Springer, Berlin, 2002. (25) R. Badea, J. A. Frey, and J. Berezovsky. Magneto-optical imaging of vortex domain deformation in pinning sites. Journal of Magnetism and Magnetic Materials, 381:463–469, 2015. 10.1016/j.jmmm.2015.01.036.

7

arXiv:1510.07059v1 [cond-mat.mes-hall] 23 Oct 2015

Supplementary Information: Mapping the domain wall pinning landscape in ferromagnetic films R. Badea and J. Berezovsky∗ Department of Physics, Case Western Reserve University, Cleveland, Ohio 44106 E-mail: [email protected]

Analysis of hysteresis loops Here we describe the procedure for extracting an effective 1D pinning potential from vortex displacement hysteresis data. An example of the analysis procedure is shown in Fig. S1. First, the raw hysteresis data is plotted in Matlab, as shown in Fig. S1a. A small linear background is subtracted from all the data, which arises from a contribution to the signal from the vortex away from the core. That is, even when the core is completely pinned, the magnetization in the surrounding region can be deformed 1 . From the measurements and simulations in Ref. 2 , we estimate the linear background to be ≈ 0.1χ0 Hs , where χ0 is the free vortex susceptibility. Next a draggable diagonal line is placed over the data which corresponds to the vortex displacement in the absence of pinning (red line in Fig. S1a). In data where the vortex is free over some range of field, this line is placed on top of these free regions. In data where the vortex occupies pinning sites over the entire range, this free line can still be determined with some accuracy as all pinning steps must begin and end on opposite sides of this line. When there is still ambiguity as to the free vortex path, the assumption that the pinning potentials are likely to be symmetric constrains ∗ To

whom correspondence should be addressed

1

a

1.5 1

(ii)

0.5

pinning features

Δy0 (arb.)

00 -0.5

(iii)

(iv)

-1

(vi)

-1.5

(v)

-2

free vortex displacement

-2.5 -3 -3.5

-4

-4000

-3

-3000

-2

-2000

-1

00

-1000

1

2

1000

2000

3

3000

4

4000

Hs (kA/m) 1

u’p (eV/nm)

u’p(y-)

-1

c

00

up (eV)

b

(i)

-5-5

A! ! C B!

0 u’p(y+)

-10 -10

-15

-3 -300

-2 -200

-1 -100

00

1 100

2 200

3 300

y (nm)

Figure S1: Steps for extracting the pinning potential from vortex displacement. a, Vortex displacement hysteresis loop as shown in main text. Red line is placed on free vortex path, blue lines are placed on pinned features. b, Points u0p (y+ ) and u0p (y− ) extracted from endpoints of blue lines in a. Blue line shows u0p constructed by connecting known points subject to conditions 1-3. c, u p from integrating u0p in b. the placement of the free path. Comparing the slope of the free vortex line to the independently measured vortex displacement vs. field allows calibration of the raw signal to vortex displacement. Next, a user-selected number of draggable (blue) lines are plotted over the raw data. These lines are placed over each pinning plateau, with the left endpoint at the point (y+ , Hx+ ) and the right endpoint at (y− , Hx− ). The lines begin at jumps on path (ii), (iii), or (vi), and end on paths (i), (iv), or (v). Each line must cross the central red line, and to ensure that the pinning sites are spatially independent, the lines must be single-valued with respect to the y-axis. (They may be, and often are, multiple-valued with respect to the H axis). It is sometimes the case that only one point of (y± , Hx± ) is visible. This can occur when the vortex occupies a particular pinning site 2

while sweeping the field in one direction, but skips over it while sweeping in the other direction. This also can occur when the vortex is pinned at the end of the scan – the higher field and position at which the vortex would escape from this pinning site are of course not measured. We estimate this point to be as symmetric as possible about the free line with respect to the other endpoint, subject to the constraint that the point occurs in a region where it is not measured (after the end of the scan, or in a jumped-over region). Uncertainty in the placement of the free vortex line and unknown pinning points are a source of the jitter seen in the pinning maps, from one 1D potential to the next. ± ± 0 As explained in the text, the measured points (y± i , Hx,i ) specify the values of u p,i (yi ) from ± ± ± 0 0 ± the condition in Eq. 1 that u0 (y± i , Hx,i ) = u0 (yi , Hx,i ) + u p,i (yi ) = 0 (because u0 (y, Hx ) is known).

Further, we know from Eq. 1 that these points are the local maxima and minima of u0p (y). The values of u0p (y± i ) are plotted as red and blue circles in Fig. S1b. Now we are free to connect the known points of u0p (y) in any way that maintains these known points as local maxima or minima. A simple way to do this would be to connect them linearly. However, it is eminently reasonable to assume that u p (−∞) = u p (∞). To enforce this with independent pinning sites, each pinning site u p,i must begin and end at the same potential (say, zero). ´ In other words, u0p,i (y)dy = 0 (condition 1). It is also reasonable to assume that the total width

over which a pinning site potential is nonzero is only a factor of 2 or so larger than the distance

+ y− i − yi (condition 2). This last assumption doesn’t affect the observed behavior much except for

the position of jumps from the free state to a pinned state, and for particularly weak pinning sites where this assumption can make the difference between a sudden jump into the pinning site or a continuous translation. Finally, for simplicity, each pinning site is defined independently, so each pinning site potential should be nonzero over a range distinct from all other pinning sites (condition 3). So we then connect u0p,i (y± i ) with the points labelled A, B and C (shown in Fig. S1b), according to the following heuristics: − 0 1. Start with B positioned so that u0p,i (y+ i ) and u p,i (yi ) are connected linearly.

3

2. If it is possible to place points A and C so as to satisfy all conditions, then do so. 3. If the conditions cannot be all satisfied, then adjust point B to the extent needed to satisfy all conditions. Figure S1b shows the constructed u0p (y) (blue line), and subsequently, integration of this piecewise linear function yields the piecewise quadratic constructed u p (y) (Fig. S1c). There are two significant caveats. First, there is sometimes substantial freedom in choosing the points A and C while still remaining consistent with the measured behavior. This can greatly affect the total depth of the potential. That is, the depth of the potential from the minimum to the known points u p (y± ) is fairly well specified because the vortex is stable over this region and its motion through this region is measured. Outside of this region however, the vortex position is typically not stable, so there is no way to know what the potential function looks like here. This problem is general to previous measurements of vortex pinning potentials, in that the extracted depth will depend on the particular functional form assumed for the pinning sites (e.g. Gaussian, Lorentzian, or piece-wise − quadratic). Second, the assumption that the pinning potential is quadratic between y+ i and yi is

equivalent to assuming that the observed pinning plateaus have constant slope with H. In some cases, the plateaus do exhibit some curvature (e.g. the highest plateau in Fig. 2a). In principle, a more accurate pinning potential could be extracted from this curvature, affecting both the shape and total depth of the pinning site. Here, we neglect this detail for simplicity. Once the pinning potential has been constructed, we can check it by simulating the expected vortex position vs. field given the constructed u p (y). To do this, we simply find a local minimum of u(y, Hx ) = u0 (y, Hx ) + u p (y) closest to an appropriate initial point, then while sweeping the AC Hx as in the experiment, follow the nearest local minimum in the “downhill” direction. The resulting simulated hysteresis loops should closely match the experimental result. Occasionally, minor ± tweaking of the (y± i , Hx,i ) points is necessary to get good agreement, though doing so typically has

very minor effects on the extracted u p (y).

4

References (1) J.A.J. Burgess, J.E. Losby, and M.R. Freeman. An analytical model for vortex core pinning in a micromagnetic disk. Journal of Magnetism and Magnetic Materials, 361:140–149, 2014. 10.1016/j.jmmm.2014.02.078. (2) R. Badea, J. A. Frey, and J. Berezovsky. Magneto-optical imaging of vortex domain deformation in pinning sites. Journal of Magnetism and Magnetic Materials, 381:463–469, 2015. 10.1016/j.jmmm.2015.01.036

5