Effect of sub-micron deformations at opposing strain rates on the micromagnetic behaviour of non-oriented electrical steel

Nature

Effect of sub-micron deformations at opposing strain rates on the micromagnetic behaviour of non-oriented electrical steel"


Play all audios:

Loading...

ABSTRACT We are entering an era of re-electrification, seeking high-power density electrical machines with minimal resource use. Significant performance gains in electrical machines have


been achieved through precise manufacturing processes, including the shaping/cutting of soft magnetic materials. However, most studies have evaluated magnetic performance at a macro level,


focusing on components, while the fundamental mechanisms, e.g., how the micromagnetic behaviour is affected by mechanical interference, remain unclear. In this study, we examine the impact


of sub-micron deformations at opposing strain rates (10−2 to 101 s−1) on the micromagnetic behaviour of soft magnetic non-oriented electrical steel. Using a diamond probe to indent within a


single grain of polycrystalline material at different velocities, we induce quasi-static and dynamic mechanical loading. Our analysis, employing magnetic force microscopy, transmission


Kikuchi diffraction, and scanning transmission electron microscopy with a pixelated detector, reveals that magnetic texture disturbances rely on the time-dependent dislocation dynamics of


the Fe-BCC material. Additionally, we compress micro-pillars to further investigate these effects under bulk-isolated deformation. These findings highlight the importance of considering even


ultra-small loads, such as nano-indentations and micro-pillar compressions, in the manufacturing of next-generation electric machines, as they can affect magnetic texture and performance.


SIMILAR CONTENT BEING VIEWED BY OTHERS STUDY ON THE INHERENT MAGNETISM AND ITS RELATIONSHIP WITH MECHANICAL PROPERTIES OF STRUCTURAL ROUND STEEL Article Open access 27 September 2022 ON THE


NATURE OF NANO TWIN-INDUCED SHEAR BAND FORMATION IN A DISPERSION STRENGTHENED COPPER ALLOY UNDER MICRO-MECHANICAL LOADING Article Open access 08 October 2024 ALLOYING EFFECTS ON DEFORMATION


INDUCED MICROSTRUCTURE EVOLUTION IN COPPER Article Open access 13 October 2024 INTRODUCTION With the global trend towards the use of cleaner energy sources and, the goal to limit the


utilisation of fossil fuels1, the expansion of electric propulsion is becoming progressively more viable2. This trend is occurring in sectors spanning society, including: ‘last-mile


logistics’ such as E-scooters3 for personal commuting; the automotive sector – driven in large part through legislative action; and the aerospace sector, where future economic security of


companies depends on their ability to embrace sustainability4. With limited resources, the importance of energy efficiency in all the aforementioned sectors is paramount for meeting future


demands and, one way to tackle this is by producing ever-more efficient (e.g., minimal magnetic losses) electrical propulsion systems. Focusing on the design and manufacture of motors,


electrical steel laminations containing Fe with a low addition of Si are widely used in both rotating (i.e., motors and generators) and static (i.e., transformers) electrical machines (Fig. 


1a). In static applications, as the magnetic flux is only required to travel in one direction, grain-oriented electrical steels are used to exploit anisotropy in the magnetic domain


behaviour. In rotating applications, non-oriented electrical steels (NOES) are used to minimise the effects of magnetocrystalline anisotropy in a bulk specimen by avoiding any strong texture


components of the microstructure. Electrical steels are used due to their relatively high permeability and low coercivity, enabling the rapid magnetisation and demagnetisation required in


rotating electrical machines5. This allows for high values of magnetic saturation, the mechanism of which depends on the spontaneous alignment of all the magnetic domains over the entire


microstructure when applying an external magnetic field. Factors which inhibit the material from reaching saturation under an applied field may be referred to as hysteresis losses (Fig. 1c),


leading to an increased energy demand to account for the inefficiencies. It is common practice to evaluate the performance of NOES, and inherently the energy losses, by their magnetisation


curves (Fig. 1c)6 (also referred to as B_H_ curves where B is the induced magnetisation and _H_ is the applied magnetic field strength). Nevertheless, these performances are not only linked


with the composition of the tested material7 but also how the magnetic domains are aligned with the crystallographic ‘easy’ axis (the preferential axis which is less energetically demanding


for the magnetic flux to traverse, e.g., < 100 > for body centred cubic (BCC) Fe within their constituent grains8,9). Studies on magnetic domains have been carried out at the


microstructural level historically with Kerr microscopy10, and more recently by forward scatter detector (FSD) imaging using an electron backscatter diffraction (EBSD) camera11. While these


more in-depth analyses have been done on as-supplied NOES (i.e., after rolling), they do not explain the intimate relationship between the follow-up manufacturing methods (Fig. 1b) of raw


material to become components in electrical motors – processes which inherently are likely to induce further losses12. Therefore, without deep knowledge of material manufacture, the


understanding of the captured magnetic domain observations is hindered. We see the effects of the convoluted magnetic domain textures, but we cannot fully understand the causes of these


disturbances, which will impair the development of more efficient electrical machines. The perturbances in the magnetic domain textures within a studied material or component can arise from


various sources, e.g., preferential heat dissipation, crystallographic texture, possible inclusions, and mechanical loading13. As most of the production steps to manufacture electrical


machines involve mechanical loading at different strain rates (e.g., cutting by various methods – punching, assembly, handling)14,15,16 there is a need to explore and understand the core


phenomena on how the associated plastic deformations can influence the magnetic domains at the microstructural level (e.g., single grain) - Fig. 1d. Prior work has also demonstrated the


influence of plastic strain on the magnetic domain texture of nanocrystalline Ni thin films through Kerr microscopy and magnetic force microscopy17, yet the strain-rate sensitivity of the


magnetic response within a single crystal orientation should also be investigated. Considering the above, it could be the case that even unexpected steps occurring in transient operations


(e.g., pick-and-place) during the manufacture of electrical machines, could affect the micromagnetic structure of NOES. A method to address these scenarios is to utilise micromechanical test


methods (e.g., nano-indentation and micro-pillar compression) to understand the strain-rate sensitivity of a NOES, with an additional focus on the magnetic domain response. Whilst B-H


testing is an acceptable method of characterising the magnetic behaviour of NOES after processing, from the practical perspective of an electrical machine designer, this macro-level approach


does little to explain the sources of variation in magnetic performances when employing different manufacturing methods that often involve complex mechanical loading. As such, inferences


can be made about what occurs to the magnetic domain texture at the micro level when (dynamic) mechanical loading is applied (e.g., impingement of garnet abrasive during abrasive waterjet


machining18,19). However, in such cases, the material deformation is uncontrolled, so definitive statements regarding the effects of mechanically induced deformation upon the magnetic


texture are difficult to interpret. In this work, nano-indentation and micro-pillar compression, which are typically used to measure micromechanical properties in stress-affected areas20,21,


are employed as a tool for inducing the stresses directly to investigate changes in the micromagnetic structure. This study of the relationship between the application of micro-loads and


the corresponding magnetic domain disturbances gives insight into the governing mechanisms which lead to increased electrical losses in NOES (Fig. 1d). RESULTS APPLICATION OF CONTROLLED


MECHANICAL LOADING Using a commercially available non-oriented electrical steel (Arnon 7 – FeSi 3 wt% Si), we induced mechanical deformation (see Fig. 1d) in single grains with their normal


axes along the < 111 > direction using a nano-indenter with a spherical diamond tip, allowing us to investigate the fundamental effects caused by a single loading scenario on the


magnetic domain structure. The < 111 > oriented grain was selected from an initial EBSD map prior to loading the sample in the nano-indenter assembly (see Supplementary Fig. 1). The


< 111 > direction was chosen to increase the proportion of magnetic domains that are oriented out of the plane at the free-surface (||XY), therefore, increasing the magnetic signal


measurement in the analyses of this surface. Whilst conventional nano-indentation is usually employed for mechanical property measurement at low strain rates22 to achieve stable,


quasi-static loading, we were particularly interested in the time-dependent behaviour of the material at the micro-level and its effects on the magnetic domain texture; this is relevant for


understanding the variations of macro-magnetic performances of NOES steels when processed by different methods. Further to nano-indentation with a spherical tip, where the bulk material


surrounding the deformed region contributes to the system through an opposing force, we also consider the scenario where the deformation is isolated from the bulk material – through


micro-pillar compression. In this study, we selected two contrasting strain rates to apply the load as 10−2 s−1 and 101 s−1, which we term LSR (low strain rate) and HSR (high strain rate),


respectively, performing indents to a depth of 0.5 μm. The strain-rate values were chosen to encompass both quasi-static loading (LSR) and dynamic loading (HSR), whilst remaining within the


data acquisition resolution of the nano-indenter load cell. The dynamic loading investigation has particular importance when considering the nature of the deformation events which may occur


during material processing and operation (e.g., pick-and-place procedures, lamination stamping, debris impact) with their inherent effects on the magnetic domain texture. MAGNETICALLY


AFFECTED ZONE AROUND SINGLE DEFORMATION SITES The investigatory procedure is detailed in Fig. 1d and the Methods section. It is already known that the complex behaviours of magnetic domains


within a ferromagnetic material are related to magnetic anisotropies that can manifest in various forms (e.g., shape, magnetocrystalline, and stress-induced anisotropies)23,24. In our case,


owing to the selection of a single unstrained grain (< 111 >), we can consider the magnetic anisotropy of a pristine, i.e., non-indented, zone to be minimal when referenced to the same


individual grain. To verify this assumption, we used an atomic force microscope (AFM) with a magnetically polarised tip to perform magnetic force microscopy (MFM) within single grains. The


selection used an initial EBSD inverse pole figure (IPF) map to identify the regions where indents were to be performed, and the pre-indentation regions were subsequently mapped using MFM


(Fig. 2a, d) on the sample free-surface (||XY) (an example of the grain selection method through EBSD and MFM is provided in Supplementary Fig. 1). After that, we indented with the


prescribed strain-rates in the pre-mapped regions and the MFM mapping procedure was repeated (Fig. 2b, e) (the loading curves for both LSR and HSR spherical nano-indentations are provided in


the Supplementary Fig. 2). The MFM mapping pre- and post-indentation enabled the evaluation of the magnetic domain texture changes, which can be directly attributed to the application of


quasi-static and dynamic mechanical loading. The MFM maps plot the centred phase shift (_θ_) in the probe oscillation measured during the lifted scan. The centred phase shift (after


subtraction of the mean) is used instead of the absolute phase shift to account for any differences in the individual cantilever harmonics or tip magnetisation which may arise due to


degradation of the consumable MFM tips. For both LSR and HSR indent locations, there is a visual difference in the magnetic domain structure before and after indentation – before, the MFM


phase distribution shows less deviation from the mean in both the positive and negative directions (line scan insets of Fig. 2a, d). After indentation, the MFM phase distribution becomes


more polarised – deviating from the mean and having greater magnitudes of phase shift between extremum (line scan insets of Fig. 2b, e), which indicates that there may be an obstruction to


magnetisation through domain motion/growth, favouring the more, energetically demanding domain rotation25. However, extracting further information visually about the magnetic domain textures


from MFM has limitations. The dimensionality reduction which occurs when scanning the MFM probe through the surface magnetic field hinders the quantitative assessment of the complex


patterns generated by the magnetic domains in the MFM image. To address this, we propose a magneto-metrological procedure to measure the radial extent of the magnetically affected region


surrounding the indentations. In Fig. 2g, h, we present a schematic representation of our procedure where we separated the MFM image into annular data subsets with regions, AR of different


radii (_R_ = 5,…, 25 μm) centred on the indent footprint. The surface morphology contribution to the MFM map was controlled by using an interleave scan, where the topographic data was


collected during the forward scan for both LSR and HSR surfaces (Fig. 2c, f respectively), enabling a constant lift height (100 nm) to be maintained between the tip and the surface during


the lifted MFM scan (an animation visualising the forward and interleave scans is provided in Supplementary Fig. 3 and Supplementary Movie 1). The MFM phase data at the lift height can be


considered valid for relatively flat samples, however, when larger features are encountered (i.e., an indent which could reach micrometre level depths) with non-planar surfaces, the tip


encounters interactions from multiple directions, the components of which cannot be accounted for with the lift height alone – inducing artefacts in the MFM phase images. To account for


this, when quantitively measuring the radial extent of the magnetic domain disorder, we did not consider the MFM phase values within the indent footprint (represented by the red circular


area in Fig. 2g) – the circular area with a diameter of 8 µm from the indentation centre (shown in Fig. 2c, f). Therefore, we evaluated the changes in magnetic texture in the regions


surrounding the indent footprint. For each AR, we determined the mean areal MFM phase shift of the surface _Mθ__A_: $${M\theta }_{A}=\frac{1}{{A}_{R}}{\mathop{\int


\int}\limits_{{A}_{R}}}\,\left|\theta \left(x,y\right)\right|{\rm{d}}x{\rm{d}}y$$ (1) where _A__R_ is the area of the annular subset, and _θ_ is the MFM phase at each position (_x_,_y_). By


applying (1) to each of the annular subsets for both LSR and HSR regimes pre-, and post-indentation, we have revealed trends pertaining to the radial extent of the magnetically affected zone


(Fig. 2i). The data for both LSR and HSR plots were normalised with respect to the mean _Mθ__A_ value of their pre-indented surface, thus, enabling the comparison of the magnetically


affected zones. Firstly, as expected from initial visual observations, due to increased stress anisotropy, both LSR and HSR indentations elevated the _Mθ__A_ relative to the baseline.


However, the magnitude of the _Mθ__A_ post-LSR was greater than that post-HSR, as was the radial distance required for _Mθ__A_ to return to the baseline. In the HSR indentation, the indenter


tip impacted the surface at a greater velocity than LSR, so one may assume that for indentations of a similar (0.5 μm) depth, increased indenter velocity would cause disturbances to the


magnetic domain structure spanning a greater volume. In HSR, the deformation occurs more rapidly than in LSR, limiting the formation and organisation of dislocation structures to the region


directly below the indent26,27,28 (which we investigated later through cross-sectional studies in Fig. 3). The generated dislocations HSR lack the well-ordered pinning sites present in the


case of LSR indentation, consequently, the interactions between abrupt dislocations directly below the indent and the magnetic domains are less pronounced on the surrounding free-surface


(||XY). The limited pinning effect29 of the dislocations on the magnetic domain walls leads to relatively fewer distortions in the surface magnetic domain texture. On the other hand, it can


be inferred that the lattice under LSR conditions has sufficient time to accommodate the deformation through dislocation movements. Therefore, in the LSR indentation, the gradual plastic


deformation that is present can allow for the dislocations to be more organised than HSR, providing well-defined pinning sites for the magnetic domain walls to interact with. The strong


interaction between the magnetic domain walls and the organised dislocations obstructs their movement away from a preferential axis (< 100 >) due to an increase in stress-anisotropy –


presenting as alterations in the magnetic domain texture. These findings highlight the role of time-dependent deformation in inducing localised changes in the magnetic domain texture on the


sample-free surface (||XY). As such, the influence of ordered dislocations interfering with surface magnetic domains, leading to alterations in the material’s magnetic properties, is an


interesting finding. It suggests that even quasi-static loading could have far-reaching effects on an electrical material’s magnetic behaviour. At a broader scope, similarly, low deformation


rates may be encountered by soft magnetic materials during commercial usage (e.g., motor lamination assembly and centripetal loads in service), leading to potential difficulties in


predicting magnetic performance. To further investigate these aspects, it is imperative to analyse the material beneath the indentation (||XZ) to establish the relationship between the depth


and magnitude of induced plastic deformation influenced by changes in strain rate, and the resultant magnetic domain texture. MAGNETICALLY AFFECTED ZONE BENEATH SINGLE DEFORMATION SITES To


observe the deformation beneath the indents, lamellae cross-sectioning (in the ||XZ plane) both LSR and HSR indents from single grains of the same crystallographic orientation were extracted


using the lift-out method within a focussed ion beam scanning electron microscope (FIB-SEM) and welded to copper grids for transmission analyses (Fig. 3a, b respectively, further details on


the spherical nano-indentation cross-sectioning procedure are provided in Supplementary Fig. 6). Through transmission Kikuchi diffraction (TKD) grain-referenced orientation deviation (GROD)


mapping, we observed that the greatest magnitudes of deformation occurred directly below the indentation in both LSR (Fig. 3e) and HSR (Fig. 3f) regimes. The deformation zones beneath both


indents also show defined contours between the areas of lower and higher misorientation. Under the LSR indent, these contours are more aligned with the loading direction (||Z) (Fig. 3e),


whereas in HSR, there is a contour which traverses the lamella (||X) (Fig. 3f). To understand the effects that mechanically induced deformation at low and high strain rates has on the


magnetic texture, we performed scanning transmission electron microscopy – differential phase contrast (STEM-DPC) (Fig. 3g, h) with a pixellated detector to map the magnitude of the in-plane


magnetic deflection, βL, beneath the indentations. In LSR indentation, plastic deformation occurs primarily along the primary slip systems, which are easily activated under quasi-static


conditions30. As the grain was loaded in the < 111 > direction, the {110} family of slip planes (the primary slip planes in BCC crystals) were easily activated. The more vertically


oriented bands observed in the DPC map (Fig. 3g) suggest a correlation between the direction of the deformation network and the arrangement of magnetic structures. Conversely, in HSR


indentation, the deformation occurs rapidly, allowing only a limited time for dislocations to move and rearrange. This results in an altered arrangement of dislocations compared to LSR


indentation. While the Peierls stress, which determines the resistance to dislocation motion, remains constant (i.e., is independent of strain rate), the rapid nature of HSR indentation


increases dislocation interactions and density – especially near the indentation surface, due to constrained mobility. This results in localised stress concentrations and limits the volume


for plastic deformation to diffuse into, leading to directionally altered magnetic structures, as seen in the two-dimensional cross-section view (Fig. 3h). Despite the region immediately


below the indentation surface having a greater magnitude of deformation than that which occurred deeper, the corresponding DPC map (Fig. 3h) highlights that there is a stronger relationship


between the dislocation network’s direction and the magnetic alignment than with the deformation magnitude. The differences in deformation behaviour between HSR and LSR highlight the


time-dependent nature of dislocation motion and interactions during rapid loading, significantly affecting both the mechanical and magnetic properties of the material. Tying information on


the magnetic domains in the cross-sectional (||XZ) observations presented in Fig. 3 with those at the free-surface (||XY), shown in Fig. 2, we can conclude that the strain rate of applied


deformation has a significant, time-dependent effect on the ability of the material to accommodate stresses. This has a marked effect on the magnetic textures at the level of a single grain.


Application of LSR loads was found to induce deformation, which primarily altered the magnetic domain texture at the free surface (||XY) – Fig. 2i (see Post-LSR line). This may be


attributed to the orientation of the activated, primary slip system intersecting with the surface, allowing the domain pinning effects to be observed on the external surface. However, for


HSR load application, the rapid deformation results in a different dislocation arrangement rather than the activation of additional slip systems as the temperature remains constant for both


LSR and HSR indentation. In this case, the deformation boundaries, with which magnetic structures were observed to align, are situated deeper below the free-surface (||XY) (shown in the


cross-section - Fig. 3h). This had the effect of an apparent reduction in _Mθ__A_ on the free-surface (||XY), requiring cross-sectional observation to determine the arrangement of the


magnetic structures – Fig. 2i (see Post-HSR line). To provide more evidence for the observations discussed around the changes of magnetic domains within plastically deformed regions, we have


performed further tests in different loading conditions. To simplify the loading scenario, and to avoid a complex stress distribution akin to that which occurred in the conventional


spherical indentation (as above), we compressed a set of micro-pillars (Fig. 4a.i, b.i – pre-compression, Fig. 4a.ii, b.ii – post-compression at LSR and HSR respectively) which were located


within a single grain oriented with the < 100 > direction of the crystal parallel to the defined X, Y, and Z axes (further details on grain identification and micro-pillar manufacture


are provided in Supplementary Fig. 4). Compressing pillars ensures that the loaded area of the grain, hence the magnetic texture, is isolated from the neighbouring material. The pillars were


deformed plastically to the same strain (_ε_ = 0.078) at opposing strain rates. Like the spherical indents, we compressed the pillars plastically at both LSR (10−2 s−1) and HSR (101 s−1) to


isolate and provide more evidence of how the different deformation mechanisms can influence the distribution of magnetic textures. To further investigate how different deformation


mechanisms influence magnetic textures, we conducted LSR and HSR compression tests on micro-pillars with the compression axis along the < 100 > direction. This orientation is different


from the < 111 > loading direction used in the previously discussed spherical indentations. The change in crystal orientation allowed us to explore how the crystallographic direction


of the applied stress affects the deformation and subsequent magnetic response. Due to this difference in orientation, we anticipated variations in the deformation patterns observed from TKD


and the resulting magnetic alignment. This approach helps isolate the effects of strain rate from those of crystal orientation, providing a clearer understanding of how deformation dynamics


influence magnetic properties. For both LSR (Fig. 4e) and HSR (Fig. 4f) pillar compression, the deformation observed in the TKD maps predominantly aligns with the primary slip planes along


the < 111 > direction. The corresponding DPC map for LSR pillar compression (Fig. 4g) shows a weakly banded magnetic structure, indicating that plastic deformation accumulates and


organises along the easily activated primary slip planes, as expected under quasi-static conditions. However, the rapid deformation in HSR pillar compression leads to a denser and more


complex dislocation network (rather than the activation of secondary slip systems) near the surface. The limited time for dislocations to move and reorganise under HSR conditions results in


disordered interactions along dislocations, which in turn create complicated stress fields. This increased dislocation density and complexity provide additional boundaries and pathways for


magnetic alignment, as seen in (Fig. 4h). DISCUSSION As the trend for electrification in society increases, it is vital to have a thorough understanding into how the functional properties of


soft magnetic materials (e.g., NOES) may be altered from material synthesis, through device manufacture, and to end usage. We demonstrate that by mechanically inducing deformation at low


(10−2 s−1) and high (101 s−1) strain rates within a single grain, the magnetic domain textures can be significantly altered. These changes are primarily due to the time-dependent dislocation


dynamics and the resulting stress fields, which introduce pinning sites that can hinder magnetic domain wall motion during subsequent magnetisation. Here, we studied these phenomena on a


commercial grade of NOES, which is composed of FeSi 3 wt% Si (with a Fe-BCC crystalline structure) as an example. We have demonstrated how, at low strain rates, plastic deformation can


influence the magnetic domain texture at the free surface (||XY) of the grain yet has little effect below the deformed site (||XZ). This could be due to the activation of primary slip


systems in quasi-static conditions, which allows plastic deformation to accumulate along the activated planes in an organised manner, as there is sufficient time for the dislocations to be


mobile. In contrast, the deformation at higher strain rates is more rapid, which leads to a more complex dislocation network, resulting in increased dislocation density and interactions near


the free surface (||XY). This creates additional boundaries and pathways for magnetic structures, affecting the alignment and behaviour of magnetic domains, therefore, there can be less


visible preferential alignment in the 2D cross-section view. Furthermore, our results illustrate how the direction of applied loading to the crystal structure leads to an


orientation-dependent response when observing changes to the magnetic texture. In our study, this was achieved by loading parallel to the < 111 > and < 100 > directions for the


bulk and bulk-isolated tests, respectively. The techniques adopted in our study (MFM, TKD, 4D-STEM) show a route to explore the micro-level magnetic properties of soft magnetic, electrical


steels when subjected to various strain rates mirroring real-life manufacturing and operational conditions, demonstrating practical relevance. As such, there is future scope for advanced


electrical material analyses at the micro level to understand how mechanical loads could impact the efficiency of electrical steels. METHODS SAMPLE PREPARATION AND SITE SELECTION FOR


MAGNETIC DOMAIN MAPPING OF THE NON-INDENTED, FREE-SURFACE (||XY) To understand the influence of indentation strain rate upon the magnetic domain texture, there is a need to start from a


surface with minimal defects (e.g., scratches, surface roughness, and contamination). As such, prior to performing the indentations, a flat sample of a commercially available NOES (Arnon 7 –


FeSi 3 wt% Si) was prepared by grinding and polishing with 0.04 µm colloidal silica. A large-scale EBSD survey of the sample was completed to build a dataset containing the crystallographic


orientation information of many grains; this is of importance as the indentations were performed in similarly oriented grains, i.e., < 111 >. See an example of large-area EBSD mapping


before indentation in Supplementary Fig. 1. MFM was used to investigate the magnetic domain texture of the free-surface (||XY) pre- and post-indentation. MFM is a modified mode of AFM where


an oscillating probe (nominal frequency = 75 KHz) with a sharp (maximum tip radius = 50 nm), magnetised (coercivity = 400 Oe) tip scans the topography of the sample surface along a single


line. This topography is then retraced by the oscillating probe at a prescribed lift height (100 nm) above the surface, and the phase shift of the oscillation along the same line is plotted.


The phase shift occurs due to interactions with the probe and the stray magnetic field above the sample surface, which is dictated by the arrangement of the magnetic domains31. The sampling


locations for MFM were chosen with the aid of an optical microscope installed on the MFM scan head and the mild etching effect, which reveals the Fe-BCC grain boundaries when polishing with


colloidal silica. Large area (50 x 50 µm) MFM scans were captured with 10% overlap between frames to generate a stitched MFM dataset. By combining MFM and EBSD datasets (an example is shown


in Supplementary Fig. 1), we were able to observe the magnetic domain texture relative to the crystallographic orientations at known positions on the sample, and grains were selected for


the initial indentation. Based on the EBSD dataset (inverse pole figure (IPF) for the Z-direction), the selected grains for the LSR and HSR indentations and subsequent further analyses have


a < 111 > axis direction parallel with the sample normal (_Z_-axis), the knowledge of which allows for the indentations to be made along the same directions in comparable grains.


NANO-INDENTATIONS (BULK DEFORMATION OF SINGLE GRAIN SITE) AT OPPOSING STRAIN RATES VS. MAPS OF MAGNETIC DOMAIN TEXTURE After the target grains were identified and mapped in the


pre-indentation state with MFM, the influence of the mechanical deformation upon the changes to the magnetic domain texture within a single grain could be investigated. A diamond spherical


indenter (Synton-MDP) with a 20 µm tip radius was used with the Alemnis Standard Assembly nano-indenter in displacement control mode (depth = 0.5 µm). This enabled the full indentation


footprint (~ 8 µm diameter due to the elastic recovery of the material and compliance in the indenter system) to be captured in both the free surface (||XY) and cross-section (||XZ) analyses


(i.e., when preparing the lamellae for TEM analysis). Two contrasting strain rates were tested to span 3 orders of magnitude (from 10−2 s−1 to 101 s−1) to understand the strain-rate


dependence of the magnetic domain texture. While the LSR and HSR indentations were made to the same depth, and therefore had comparable geometry, the effect of dynamic indentation would


inevitably lead to variations in lattice deformation and misorientation. Therefore, the application of contrasting strain rates would cause the magnetic domains to distort through different


mechanisms. To control interference from environmental variables, the indentations were completed at a high vacuum inside the chamber of an FEI Quanta650-ESEM, and the drift in the piezo


actuator was stabilised by maintaining the nano-indenter assembly at these conditions for over 12 h. MAPPING OF MAGNETIC DOMAIN DISTORTION INDUCED BY NANO-DEFORMATION WITHIN A SINGLE GRAIN:


FREE-SURFACE (||XY) Post-indentation, the MFM maps were repeated at the indent sites, enabling the analysis of how the magnetic domain texture becomes distorted through the introduction of


deformation alone. The degree to which the magnetic domain distortion extends beyond the physical boundaries of the indentation was enabled by the superposition of the traditional AFM


topography data (collected on a second data channel during the MFM scan) over the MFM data. The topography data was levelled using a least squares plane fit, sampling the entire scan area


but omitting features above (e.g., indentation pile-up) and below (i.e., the indentation volume) the surface. The overlapping of AFM and MFM mapping facilitated the spatial interpretation of


data. MICRO-PILLAR COMPRESSION (ISOLATED FROM BULK DEFORMATION IN A SINGLE GRAIN) AT OPPOSING STRAIN RATES FOR MAPPING THE MAGNETIC DOMAIN TEXTURE To select the appropriate level of applied


strain, we compressed two reference micro-pillars at low and high strain rates (LSR and HSR) to failure and extracted the stress-strain curves. For this calibration data to be relevant for


the subsequent compressed pillars, we ensured that the crystallographic orientation of each pillar at a site within a single grain was identical by completing a large EBSD map prior to


pillar milling (Supplementary Fig. 4a). A ZEISS Crossbeam 550 FIB-SEM fitted with an OINA Symmetry EBSD detector was used for this purpose. A large grain (area = 53,358 µm2 shown in


Supplementary Fig. 4b.i) was identified for pillar milling (Fig. 5a) with orientation (< 100 > ||Z) so that it can accommodate many pillars while avoiding the errors that might be


introduced by grain boundaries. For this, markers (X, +, K) (annotated in Fig. 5a) were milled into the surface of the chosen grain to demarcate its boundaries so that the testing site could


be located in subsequent operations. Two pads of Pt (75 μm x 75 μm) were deposited within the grain using the gas injection system (GIS) within the FIB to protect the material against Ga +


implantation. A local reference grid (e.g., D, E, F and 1, 2, 3) was also milled adjacent to the Pt protective pads for pillar identification (Fig. 5b). The pillars were milled to a nominal


diameter of 5 μm and height of 10 μm for use with the diamond indenter (6 μm diameter flat punch). The trenches surrounding the pillars were set to a diameter of 25 μm to aid in measuring


the actual pillar heights and to provide probe access when removing the pillars later. For pillar milling, a roughing pass at 7 nA was performed, followed by two subsequent finishing passes


of 1.5 nA and 0.3 nA. The FIB apertures were aligned to a 50 pA reference probe prior to the milling process to reduce exposure to higher FIB currents (i.e., reducing Ga + implantation) and


to ensure concentricity between passes. The uncompressed, calibration micro-pillars were imaged at high resolution on a FEG-SEM to determine their diameter, height, and Pt layer thickness


prior to indentation. The calibration pillars were compressed to failure at low and high strain rates to establish reference stress-strain curves (Fig. 5c, d, respectively). From the


acquired curves, we selected the values of strain required to enter the plastic region (\(\varepsilon \cong\) 0.078) (in situ videos (Supplementary Movies 2–5) show the micro-pillar


compression and, are detailed in Supplementary Fig. 5). Following pillar compression, high-resolution FEG-SEM imaging was completed to identify failure mode, shearing direction and real


plastic strain for the cases where the test was stopped prior to slipping/shearing activity. Using this calibration data, we compressed pillars located in a single grain to a level of strain


prior to failure; this enabled the evaluation of changes in the magnetic texture after deformation at different strain rates. ANALYSIS OF THE STRAIN-RATE SENSITIVITY OF THE MICROSTRUCTURE


BENEATH BOTH BULK AND BULK-ISOLATED DEFORMATION SITES WITHIN SINGLE GRAIN AND THE CORRELATION WITH MAGNETIC TEXTURE The following investigatory method applies to both nano-indentations


(bulk) and micro-pillars (bulk-isolated). As the strain rate was varied, residual plastic deformation and elastic recovery of different magnitudes occurred beneath the free surface (||XY),


requiring the investigation of a cross-sectional view (||XZ) of the material under both the indentations and the compressed pillars. Using the FEI Quanta200-ESEM, a protective Pt layer


approximately 1 µm thick was deposited over the indentations and compressed pillars using the FIB. A low ion beam current of 50 pA was used to protect the free surface (||XY) against


amorphization and ion implantation during the subsequent milling operations. The cross-sectional FIB milling was then completed at an ion beam current of 3 nA, the lamellae were lifted out,


and Pt welded to a copper lift-out grid. Thinning of the lamellae was performed on the Zeiss Crossbeam 550 FIB-SEM to achieve a thickness of approximately 150 nm – enabling further


transmission studies (further details regarding the lamella preparation method for both the spherical nano-indentations and the micro-pillars are provided in Supplementary Figs. 6 and 7,


respectively). To map the relative sub-granular deformation occurring beneath the indentations and compressed pillars, off-axis TKD was performed with an accelerating voltage of 30 kV, an


electron beam current of 1.0 nA, and a pixel size of 0.02 µm. These parameters allowed us to obtain a high level of angular precision for mapping the grain-referenced orientation deviation


(GROD) angle for each scan point in the crystal lattice beneath the indentations and compressed pillars. It should also be noted that the act of thinning the deformed regions (indentations


and compressed pillars) into lamellae may cause some dislocation relaxation – which could lead to an apparent reduction in the dislocation density when measured with TKD. A solution to this


problem has been proposed and explored successfully by Kalácska et al.32 where serial FIB slicing was used on deformed, square, copper micro-pillars to observe the dislocation network in 3D


to obtain more reliable results on the geometrically necessary dislocation (GND) density. However, this would require the grain of interest to be positioned on the edge of the sample, and


the method is destructive – limiting further analysis of the sample with STEM methods, placing this method outside the scope of the current study. Scanning transmission electron microscopy


(STEM), was used for the isolation of contrast arising from magnetic origins. The STEM images were captured using a Gatan OneView pixelated detector within a JEOL 2100 + TEM at 200 kV (Fig. 


6). A large camera length (150 cm) was used with a 10 µm aperture, giving a convergence semi-angle of 2.35 mrad, enabling the imaging of the central transmitted disk in the convergent beam


electron diffraction (CBED) image. It should be noted that greater angular precision of the central disk could be achieved by using a larger aperture (e.g., using a 50 µm aperture would


yield a convergence semi-angle of 13.2 mrad); however, this would limit the working distance of the electron beam causing the focus value to vary between the upper and lower faces of the


sample. For both the LSR and HSR lamellae, scans measuring ~ 6 x 4 µm were made in the same locations as the TKD maps. To process the 4D-STEM33 data, the CBED images were extracted from the


native.dm4 files using the HyperSpy library34. It should be noted that file size becomes a significant constraint with 4D-STEM datasets, so the detector may be binned at the data collection


stage to ensure that the dataset is not too large for practical post-processing. To reduce the dataset size further, we computed an average CBED image from the full scan area and


square-cropped all of the CBED images to the central disk only. To correct for the influence of scan position on the central disk location as the electron beam is translated over the sample,


we applied a Sobel operator to emphasise the disk edges, fitting a circle to the edge image using Taubin’s singular value decomposition (SVD) method35. An affine transformation was applied


to the individual CBED images to position the central disk in the centre of the cropped region. Following the scan correction procedure, we calculated the centre-of-mass (CoM) of the central


spot intensity, producing a map of beam displacements after interaction with the sample. These displacements are caused by interactions with the magnetic domains within the material and the


electron beam due to the Lorentz force, with the deflection angle, _β__L_, relating to magnetic induction, BS, by: $${\beta }_{L}=-\frac{e\lambda }{h}{{{\bf{B}}}}_{{{\bf{s}}}}t,$$ (2) where


_e_ is the magnitude of electron charge, _λ_ is the relativistic electron wavelength at an accelerating voltage of 200 kV, _t_ is the sample thickness, and _h_ is Planck’s constant36,37.


Mapping _β__L_ in the same region as the TKD maps using (2) has enabled the correlative analysis of sub-surface deformation structures with magnetic textures, revealing the strain-rate


dependency of their relationship. DATA AVAILABILITY The datasets generated during and/or analysed during the current study are available from the corresponding author on request. Source data


are provided in this paper. REFERENCES * Luderer, G. et al. Residual fossil CO2 emissions in 1.5–2 °C pathways. _Nat. Clim. Change_ 8, 626–633 (2018). Article  ADS  CAS  Google Scholar  *


Bistline, J. E. T. et al. Economy-wide evaluation of CO2 and air quality impacts of electrification in the United States. _Nat. Commun._ 13, 1–12 (2022). Article  Google Scholar  * Asensio,


O. I., Apablaza, C. Z., Lawson, M. C., Chen, E. W. & Horner, S. J. Impacts of micromobility on car displacement with evidence from a natural experiment and geofencing policy. _Nat.


Energy_ 7, 1100–1108 (2022). Article  ADS  Google Scholar  * Aghahosseini, A. et al. Energy system transition pathways to meet the global electricity demand for ambitious climate targets and


cost competitiveness. _Appl. Energy_ 331, 120401 (2023). Article  Google Scholar  * Han, L. et al. A mechanically strong and ductile soft magnet with extremely low coercivity. _Nature_ 608,


310–316 (2022). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Grössinger, R., Sassik, H., Holzer, D. & Pillmayr, N. Magnetic characterization of soft magnetic


materials—experiments and analysis. _J. Magn. Magn. Mater._ 254, 7–13 (2003). Article  ADS  Google Scholar  * Hawezy, D. The influence of silicon content on physical properties of


non-oriented silicon steel. _Mater. Sci. Technol._ 33, 1560–1569 (2017). Article  ADS  CAS  Google Scholar  * Bozorth, R. M. _Ferromagnetism_. (Van Nostrand, Princeton, N.J., 1951). *


Cullity, B. D. _Introduction to Magnetic Materials_. (IEEE/Wiley, 2015). * Shilling, J. & Houze, G. Magnetic properties and domain structure in grain-oriented 3% Si-Fe. _IEEE Trans.


Magn._ 10, 195–223 (1974). * Gallaugher, M., Brodusch, N., Gauvin, R. & Chromik, R. R. Magnetic domain structure and crystallographic orientation of electrical steels revealed by a


forescatter detector and electron backscatter diffraction. _Ultramicroscopy_ 142, 40–49 (2014). Article  CAS  PubMed  Google Scholar  * Bali, M. & Muetze, A. The degradation depth of


non-grain oriented electrical steel sheets of electric machines due to mechanical and laser cutting: a state-of-the-art review. _IEEE Trans. Ind. Appl._ 55, 366–375 (2019). * Ickler, T.,


Meckbach, H., Zeismann, F. & Brückner-Foit, A. Assessing the influence of crystallographic orientation, stress and local deformation on magnetic domains using electron backscatter


diffraction and forescatter electron imaging. _Ultramicroscopy_ 198, 33–42 (2019). Article  CAS  PubMed  Google Scholar  * Bali, M. & Muetze, A. Modeling the effect of cutting on the


magnetic properties of electrical steel sheets. _IEEE Transactions on Industrial Electronics_ 64, 2547–2556 (2017). * Saleem, A., Goldbaum, D., Brodusch, N., Gauvin, R. & Chromik, R. R.


Microstructure and mechanical property connections for a punched non-oriented electrical steel lamination. _Mater. Sci. Eng. A Struct. Mater._ 725, 456–465 (2018). * Saleem, A., Alatawneh,


N., Rahman, T., Lowther, D. A. & Chromik, R. R. Effects of laser cutting on microstructure and magnetic properties of non-orientation electrical steel laminations. _IEEE Trans. Magn._


56, 1–9 (2020). * Zhang, K. et al. Ion-induced magnetic texturing of Ni films: Domain structure and strain. _Nucl. Instrum. Methods Phys. Res. B_ 257, 379–382 (2007). Article  ADS  CAS 


Google Scholar  * Schoppa, A., Louis, H., Pude, F. & Von Rad, C. Influence of abrasive waterjet cutting on the magnetic properties of non-oriented electrical steels. _J. Magn. Magn.


Mater._ 254–255, 370–372 (2003). Article  ADS  Google Scholar  * Winter, K. et al. How non-conventional machining affects the surface integrity and magnetic properties of non-oriented


electrical steel. _Mater. Des._ 210, 110051 (2021). Article  CAS  Google Scholar  * Ding, Y., Gallaugher, M., Brodusch, N., Gauvin, R. & Chromik, R. R. Effect of a coating induced


residual stress on magnetic domain structure in non-oriented electrical steels. _Microsc. Microanal._ 20, 894–895 (2014). * Uchic, M. D., Shade, P. A. & Dimiduk, D. M. Plasticity of


micrometer-scale single crystals in compression. _Annu. Rev. Mater. Res._ 39, 361–386 (2009). Article  ADS  CAS  Google Scholar  * Kubaschinski, P., Gottwalt-Baruth, A., Tetzlaff, U.,


Altenbach, H. & Waltz, M. Modelling and simulation of the hardness profile and its effect on the stress-strain behaviour of punched electrical steel sheets. _Materwiss. Werksttech._ 54,


512–526 (2023). Article  Google Scholar  * Leuning, N., Steentjes, S., Heller, M., Korte-Kerzel, S. & Hameyer, K. On the correlation of crystallographic macro-texture and magnetic


magnetization anisotropy in non-oriented electrical steel. _J. Magn. Magn. Mater_. 490, 165485 (2019). Article  CAS  Google Scholar  * Landgraf, F. J. G., Yonamine, T., Emura, M. &


Cunha, M. A. Modelling the angular dependence of magnetic properties of a fully processed non-oriented electrical steel. _J. Magn. Magn. Mater._ 254, 328–330 (2003). * Qiu, F., Klug, M. J.,


Evi Tian, G., Hu, P. & McCord, J. Influence of magnetic domain wall orientation on Barkhausen noise and magneto-mechanical behavior in electrical steel. _J. Phys. D Appl. Phys._ 52,


265001 (2019). Article  ADS  CAS  Google Scholar  * Peng, X., Wang, L., Xu, W. & Liang, Z. Crystal orientation dependence of the portevin–Le chatelier effect in instrumented indentation:


A case study in twinning-induced plasticity steels. _Metals_ 12, 439 (2022). * Saleh, M. et al. Oblique cross-section nanoindentation for determining the hardness change in ion-irradiated


steel. _Int. J. Plast._ 112, 242–256 (2019). Article  CAS  Google Scholar  * Bradby, J. E., Williams, J. S., Wong-Leung, J., Swain, M. V. & Munroe, P. Transmission electron microscopy


observation of deformation microstructure under spherical indentation in silicon. _Appl. Phys. Lett._ 77, 3749–3751 (2000). Article  ADS  CAS  Google Scholar  * Lindquist, A. K., Feinberg,


J. M., Harrison, R. J., Loudon, J. C. & Newell, A. J. Domain wall pinning and dislocations: Investigating magnetite deformed under conditions analogous to nature using transmission


electron microscopy. _J. Geophys. Res. Solid Earth_ 120, 1415–1430 (2015). Article  ADS  Google Scholar  * Weinberger, C. R., Boyce, B. L. & Battaile, C. C. Slip planes in bcc transition


metals. _Int. Mater. Rev._ 58, 296–314 (2013). Article  CAS  Google Scholar  * Passeri, D. et al. Magnetic force microscopy: Quantitative issues in biomaterials. _Biomatter_ 4, e29507


(2014). Article  PubMed  PubMed Central  Google Scholar  * Kalácska, S. et al. Investigation of geometrically necessary dislocation structures in compressed Cu micropillars by 3-dimensional


HR-EBSD. _Mater. Sci. Eng. A_ 770, 138499 (2020). Article  Google Scholar  * Ophus, C. Four-dimensional scanning transmission electron microscopy (4D-STEM): From scanning nnodiffraction to


ptychography and beyond. _Microsc. Microanal._ 25, 563–582 (2019). Article  ADS  CAS  PubMed  Google Scholar  * de la Peña, F. et al. hyperspy/hyperspy: Release v1.7.3.


https://doi.org/10.5281/zenodo.7263263 (2022). * Taubin, G. Estimation of planar curves, surfaces, and nonplanar space curves defined by implicit equations with applications to edge and


range image segmentation. _IEEE Trans. Pattern Anal. Mach. Intell._ 13, 1115–1138 (1991). Article  Google Scholar  * Krajnak, M., McGrouther, D., Maneuski, D., O’Shea, V. & McVitie, S.


Pixelated detectors and improved efficiency for magnetic imaging in STEM differential phase contrast. _Ultramicroscopy_ 165, 42–50 (2016). Article  CAS  PubMed  Google Scholar  * Chapman, J.


N. The investigation of magnetic domain structures in thin foils by electron microscopy. _J. Phys. D Appl. Phys._ 17, 623 (1984). Article  ADS  CAS  Google Scholar  Download references


ACKNOWLEDGEMENTS We would like to acknowledge the support of Dr Michael W. Fay from the Nanoscale and Microscale Research Centre at the University of Nottingham and Dr Oriol Gavalda Diaz,


for assisting with the Scanning Transmission Electron Microscopy data collection. We would like to thank the Nanoscale and Microscale Research Centre (nmRC) at the University of Nottingham


for access to the electron microscope facilities. This work was supported by the Engineering and Physical Sciences Research Council (EPSRC) under grants EP/L022494/1, EP/S021434/1,


EP/W006413/1, and the University of Nottingham. We would also like to thank Mr. Artem Bogatyrev from the Rolls-Royce UTC in Manufacturing and On-Wing Technology at the University of


Nottingham for assisting with the Scanning Transmission Electron Microscopy data analysis. This work was also supported by the EPSRC through the DTP 2018–19 University of Nottingham under


grant EP/T517902/1. This study was supported by the Nottingham research fellowship programme and Beacons of Excellence granted to Zhirong Liao. AUTHOR INFORMATION AUTHORS AND AFFILIATIONS *


Rolls-Royce University Technology Centre in Manufacturing and On-Wing Technology, Faculty of Engineering, University of Nottingham, Nottingham, UK Kieran Winter, Zhirong Liao, Erik Abbá, 


Jose A. Robles Linares & Dragos Axinte Authors * Kieran Winter View author publications You can also search for this author inPubMed Google Scholar * Zhirong Liao View author


publications You can also search for this author inPubMed Google Scholar * Erik Abbá View author publications You can also search for this author inPubMed Google Scholar * Jose A. Robles


Linares View author publications You can also search for this author inPubMed Google Scholar * Dragos Axinte View author publications You can also search for this author inPubMed Google


Scholar CONTRIBUTIONS K. Winter, Z. Liao, and D. Axinte conceived the idea. K. Winter conducted the initial experiments for, Magnetic Force Microscopy, Transmission Kikuchi Diffraction, and


Scanning Transmission Electron Microscopy for the nano-indentation samples. K. Winter, J. A. Robles Linares, and E. Abbá fabricated and compressed the micro-pillars. J. A. Robles Linares


analysed the micro-pillar compression data. K. Winter conducted the experiments for Transmission Kikuchi Diffraction and Scanning Transmission Electron Microscopy for the micro-pillar


samples and analysed the results with J. A. Robles Linares and E. Abbá. K. Winter, D. Axinte, E. Abbá, J. A. Robles Linares, and Z. Liao interpreted the data. K. Winter, D. Axinte and Z.


Liao wrote and edited the manuscript. CORRESPONDING AUTHORS Correspondence to Zhirong Liao or Dragos Axinte. ETHICS DECLARATIONS COMPETING INTERESTS The authors declare no competing


interests. PEER REVIEW PEER REVIEW INFORMATION _Nature Communications_ thanks Daniele Passeri and the other anonymous reviewer(s) for their contribution to the peer review of this work. A


peer review file is available. ADDITIONAL INFORMATION PUBLISHER’S NOTE Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.


SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION DESCRIPTION OF ADDITIONAL SUPPLEMENTARY FILES SUPPLEMENTARY MOVIE 1 SUPPLEMENTARY MOVIE 2 SUPPLEMENTARY MOVIE 3 SUPPLEMENTARY MOVIE 4


SUPPLEMENTARY MOVIE 5 SOURCE DATA SOURCE DATA TRANSPARENT PEER REVIEW FILE RIGHTS AND PERMISSIONS OPEN ACCESS This article is licensed under a Creative Commons Attribution 4.0 International


License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source,


provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons


licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by


statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit


http://creativecommons.org/licenses/by/4.0/. Reprints and permissions ABOUT THIS ARTICLE CITE THIS ARTICLE Winter, K., Liao, Z., Abbá, E. _et al._ Effect of sub-micron deformations at


opposing strain rates on the micromagnetic behaviour of non-oriented electrical steel. _Nat Commun_ 15, 9010 (2024). https://doi.org/10.1038/s41467-024-53346-7 Download citation * Received:


25 August 2023 * Accepted: 08 October 2024 * Published: 18 October 2024 * DOI: https://doi.org/10.1038/s41467-024-53346-7 SHARE THIS ARTICLE Anyone you share the following link with will be


able to read this content: Get shareable link Sorry, a shareable link is not currently available for this article. Copy to clipboard Provided by the Springer Nature SharedIt content-sharing


initiative


Trending News

Past rates: 2023 va dic rates for spouses and dependents | veterans affairs

------------------------- HOW DIC MAY AFFECT YOUR VA SURVIVORS PENSION OR SURVIVOR BENEFIT PLAN HOW DO MY PAYMENTS FOR D...

Smithsonian Physical Tables | Nature

ABSTRACT SINCE the sixth edition of this standard volume of tables was reviewed in NATURE for July 5, 1915, extensive ch...

Defects break symmetry rules | Nature Materials

Radiaton damage to graphite cores in nuclear reactors creates defects that store energy, and sometimes — more dangerousl...

6 conspiracy theories that prove osho’s death was as controversial as his life - scoopwhoop

Netflix’s _Wild Wild Country___ is absolutely blowing up. While the documentary on controversial guru Osho told the stor...

Little-known spanish city on top of huge mountain with panoramic views

PERCHED ON THE EDGE OF A CANYON, THIS HISTORIC TOWN IS STEEPED IN HISTORICAL SIGNIFICANCE AND KNOWN FOR ITS HIGH-ALTITUD...

Latests News

Effect of sub-micron deformations at opposing strain rates on the micromagnetic behaviour of non-oriented electrical steel

ABSTRACT We are entering an era of re-electrification, seeking high-power density electrical machines with minimal resou...

A world off balance | thearticle

Coronavirus has now spread, from the food markets of Wuhan to the stock markets of the world. Late on Monday afternoon, ...

Polymorphisms of the androgen transporting gene slco2b1 may influence the castration resistance of prostate cancer and the racial differences in respo

ABSTRACT BACKGROUND: Organic anion-transporting polypeptides (OATPs) encoded by SLCO mediate the cellular uptake of many...

I was deathly afraid of the sea — now i work to protect it

* WHERE I WORK * 15 May 2023 Conservationist Loyiso Dunga aims to guard kelp, a crucial resource in South African waters...

Of methods and mapping | Nature Medicine

ABSTRACT THE METHODOLOGICAL ADVANCES OF THE 1970S HAVE CONTRIBUTED TO THE PREPARATION OF A CHROMOSOMAL GENE MAP THAT SHO...

Top