You are on page 1of 17

JID: ACTBIO

ARTICLE IN PRESS [m5G;August 26, 2022;19:6]


Acta Biomaterialia xxx (xxxx) xxx

Contents lists available at ScienceDirect

Acta Biomaterialia
journal homepage: www.elsevier.com/locate/actbio

Full length article

Correlating the microstructural architecture and macrostructural


behaviour of the brain
Mayra Hoppstädter a, Denise Püllmann a, Robert Seydewitz a, Ellen Kuhl b, Markus Böl a,∗
a
Institute of Mechanics and Adaptronics, Technische Universität Braunschweig, Braunschweig D-38106, Germany
b
Departments of Mechanical Engineering and Bioengineering, Wu Tsai Neurosciences Institute, Stanford University, Stanford, CA, United States

a r t i c l e i n f o a b s t r a c t

Article history: The computational simulation of pathological conditions and surgical procedures, for example the re-
Received 15 April 2022 moval of cancerous tissue, can contribute crucially to the future of medicine. Especially for brain surgery,
Revised 2 August 2022
these methods can be important, as the ultra-soft tissue controls vital functions of the body. However,
Accepted 16 August 2022
the microstructural interactions and their effects on macroscopic material properties remain incompletely
Available online xxx
understood. Therefore, we investigated the mechanical behaviour of brain tissue under three different
Keywords: deformation modes, axial tension, compression, and semi-confined compression, in different anatomical
Microstructural analyses regions, and for varying axon orientation. In addition, we characterised the underlying microstructure in
Histological investigation terms of myelin, cells, glial cells and neuron area fraction, and density. The correlation of these quantities
Axial tension tests with the material parameters of the anisotropic Ogden model reveals a decrease in shear modulus with
Axial compression tests increasing myelin area fraction. Strikingly, the tensile shear modulus correlates positively with cell and
Semi-confined compression tests
neuronal area fraction (Spearman’s correlation coefficient of rs = 0.40 and rs = 0.33), whereas the com-
Material modelling
Sus scrofa domesticus
pressive shear modulus decreases with increasing glial cell area (rs = −0.33). Our study finds that tissue
non-linearity significantly depends on the myelin area fraction (rs = 0.47), cell density (rs = 0.41) and glial
cell area (rs = 0.49). Our results provide an important step towards understanding the micromechanical
load transfer that leads to the non-linear macromechanical behaviour of the brain.

Statement of significance

Within this article, we investigate the mechanical behaviour of brain tissue under three different defor-
mation modes, in different anatomical regions, and for varying axon orientation. Further, we characterise
the underlying microstructure in terms of various constituents. The correlation of these quantities with
the material parameters of the anisotropic Ogden model reveals a decrease in shear modulus with in-
creasing myelin area fraction. Strikingly, the tensile shear modulus correlates positively with cell and
neuronal area fraction, whereas the compressive shear modulus decreases with increasing glial cell area.
Our study finds that tissue non-linearity significantly depends on the myelin area fraction, cell density,
and glial cell area. Our results provide an important step towards understanding the micromechanical
load transfer that leads to the non-linear macromechanical behaviour of the brain.
© 2022 The Author(s). Published by Elsevier Ltd on behalf of Acta Materialia Inc.
This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

1. Introduction and meninges. Abnormal loads can only occur under pathological
conditions, e.g., due to traumatic brain injuries [1–3] or diseases
The brain to be part of the central nervous system is located such as cancer [4], the latter often requiring surgical treatment to
in the cranial cavity, where under physiological conditions it is remove the tumor cells. Another example of such a procedure is
well protected from mechanical stress by the cerebrospinal fluid decompressive craniectomy, which is used to relieve the pressure
caused by brain swelling [5]. The study by Weickenmeier et al.
[6] showed that a patient-specific simulation can be used to plan

Corresponding author. and optimise the performance of a decompressive craniectomy. To
E-mail address: m.boel@tu-braunschweig.de (M. Böl).

https://doi.org/10.1016/j.actbio.2022.08.034
1742-7061/© 2022 The Author(s). Published by Elsevier Ltd on behalf of Acta Materialia Inc. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

Please cite this article as: M. Hoppstädter, D. Püllmann, R. Seydewitz et al., Correlating the microstructural architecture and macrostruc-
tural behaviour of the brain, Acta Biomaterialia, https://doi.org/10.1016/j.actbio.2022.08.034
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

use the potential of computer simulation in medicine, it crucial to understanding of the load transfer mechanisms in the brain and
understand the mechanisms of load transfer in brain tissue. With will serve as a basis for the development of suitable microstruc-
this knowledge, computer models of the brain can be created and tural models in the future.
validated [7–10]. To gain this knowledge, a mechanical, experimen-
tal characterisation of the brain tissue is required. 2. Anatomy and microstructure of the brain
A large number of in vitro experiments to characterise brain tis-
sue have been carried out in the past. Besides shear experiments, The brain is the control organ of the human body. It consists of
which can be divided into oscillatory [11–17] and simple shear a three-dimensional network of about 86 billion neurons responsi-
[18–26], most studies are based on uniaxial tensile [23,26–33], ble for receiving and processing information [73]. The largest pro-
compression [19,20,23,33–45], and indentation tests [21,38,46–54]. cessing area is the cerebral cortex, which forms the outer layer
Thereby tissue from humans [23,26,28,55], monkeys [56,57], cattles of the cerebrum and includes the characteristic convolutions and
[13,33,49,50], pigs [11,12,29,34], and rodents [32,52,58–60] were folds that allow for a relatively large surface area in the limited
used. Due to the different species and the fact that most studies space of the cranial cavity, see Fig. 1. It is connected to subcorti-
have developed different testing protocols, the results of the me- cal neuronal regions such as the basal ganglia, which transmit sig-
chanical tests are difficult to compare. An example of this is the nals to and from the cortex. Transmission, in turn, occurs through
determination of pre-load. Depending on the definition, there are axons running through connecting regions such as the corona ra-
large differences in the maximum stress/strain. Since the extreme diata and the corpus callosum. The former connects subcortical
softness of brain tissue implies very small forces, many studies regions to the cortex, the latter neurons from both hemispheres.
point to an optically determined contact point as the origin of Based on this functional difference and macroscopic appearance,
the force. However, this depends on many factors, which will be the brain is divided into grey and white matter. Fig. 1 illustrates a
discussed throughout this publication. Nevertheless, incredible in- coronal section of the porcine brain with the microstructural com-
sights into the mechanical properties of the brain have been gained positions of the above anatomical regions visualised by Klüver–
in the last 60 years [15,61]. Most studies agree that brain tissue de- Barrera staining. The cortex, representing the grey matter, consists
pends on the loading rate [23,27,34,40,42,44]. Thus, an increase in mainly of neurons and the second important cell type within the
the loading rate is associated with an increase of the reaction force. brain, the glial cells, also called neuroglia. Neuroglia are divided
Further, the behaviour under tension and compression differs in into microglia and macroglia, which include astrocytes and oligo-
form and extent, which is also referred to as tension-compression dendrocytes. The former take on stabilising and metabolic tasks,
asymmetry (TCA) [23,26,27,31,33,62]. In addition, there are regional the latter form a fatty sheath, called myelin, around the axons,
dependencies [15,19,23,26,29,32,59], as the brain can be divided which enables a high transmission speed of impulses [74]. In the
into grey and white matter from a macroscopic point of view. staining, both cell types, neurons and glia, are highlighted in pur-
The reason for this division is the underlying microstructure, see ple and differ in shape and size. The white matter is represented
Section 2. The heterogeneity of this microstructure thus leads to a by the corpus callosum and the corona radiata. Here, myelinated
change in the material properties of the brain tissue. axons and oligodendrocytes are arranged predominately along the
However, there is a lack of understanding how exactly the mi- axons. This is evident from the stained samples, as the blue-stained
crostructure influences the macroscopic behaviour. It is not clear, myelin serves as an indirect visualisation of the axons. In addition,
e.g., whether and which cells have an influence on the nonlinear the purple stained glial cells line up in the direction of the myeli-
material behaviour of brain tissue. To characterise this influence, nated axons. In contrast, the basal ganglia are present in both grey
there are micro-level studies in which individual components of and white matter and contain both characteristic microstructures.
the microstructure are mechanically tested [63–69], and macrome- It is worth noting that the brain is densely packed with the cell
chanical studies in which histological data are used to establish network and the ECM therefore occupies only 10–20% of the total
a correlation between microstructural and mechanical properties brain volume [75].
[32,50,60,70–72]. In particular, a positive correlation was found be-
tween axon volume fraction and white matter stiffness [63,64]. In
addition, Javid et al. [66] reported up to three times higher ini- 3. Materials and methods
tial and long-term elastic modulus of axons compared to the ex-
tracellular matrix (ECM) in the brainstem. Studies on the mechan- The study was exempted from ethical committee review accord-
ical properties of individual neurons yielded contrasting results: ing to national regulations (German Animal Welfare Act), as brains
Bernick et al. [65] reported rate-dependent behaviour, while An- from healthy, domestic pigs were obtained from a slaughterhouse
thonisen et al. [69] found no statistically significant dependence. immediately after animal sacrifice.
However, these micro-level approaches are only capable of captur-
ing the mechanical properties of a single component. Studies at the 3.1. Tissue sample dissection and processing
macro level, on the other hand, take into account the interaction
of the various microstructural components. E.g., Antonovaite et al. Due to the limited availability of fresh human tissue and the
[60] used atomic force microscopy on mouse brains in combination anatomical similarity between human and pig brains [76,77], ex-
with immunohistochemical staining, and found a positive correla- periments were performed on seventeen (n = 17) procine brains
tion between myelin and astrocyte content and stiffness. Further- (sus scrofa domestica) aged 7.2 ± 1.8 months and weighing
more, a negative correlation between general cell density, without 110 ± 10 g. To ensure damage-free transport, the brains were
subdivision into the different cell types, and the shear modulus left in the skull and transported to the laboratory within a max-
was documented in Budday et al. [72]. However, contradictory re- imum of 1 h after animal sacrifice. In the laboratory, the brains
sults such as the correlation of cell density with stiffness, which were removed, weighed, measured, and stored at 4°C for a few
is both positive [32,71] and negative [70], suggest that a general minutes until the samples were cut. During all processing steps,
statement about the correlation of microstructure with mechanical the samples were continuously hydrated with Dulbecco’s phos-
behaviour is not yet possible. phate buffered saline (DPBS). The samples for the mechanical tests
Therefore, the aim of this work is to investigate the origin of were taken from four anatomical regions, see Fig. 1: The gray mat-
these discrepancies by relating microstructural components to me- ter cortex (CX), the corpus callosum (CC), and the corona radiata
chanical properties. We expect that this study will lead to a better (CR) of the white matter, as well the basal ganglia (BG), a mixed

2
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Fig. 1. Coronal section of the brain with highlighted regions from which samples were taken in the present study. The histological sections schematically show the mi-
crostructure of the brain and its components (single-arrow: neuron, double-arrow: glia cell, triple-arrows: myelin). Green lines illustrate the axon orientations. Scale bars;
black: 110 μm, white: 10 mm.

Table 1
Summary of one hundred and ninety experiments depending on the dissection region (CX, BG, CC, CR),
orientation (0°, 45°, 90°, I, II, III) and deformation state (uniaxial tension/compression, semi-confined com-
pression).

Region Orientation [°] Uniaxial tension Uniaxial compression Semi-confined compression

CX – – 14 –
BG – 18 19 –
CR 0 14 15 –
90 – 15 –
CC 0 18 14 –
45 – 6 –
90 7 15 –
I – – 6
II – – 10
III – – 10
Total 57 98 26

area of gray and white matter. For further testing, isotropic mate- at a frequency of 5 Hz and recording the transverse deformation
rial characteristics are assumed for the grey matter [78] and a po- of the specimen. This camera was used to assess the correct per-
tentially anisotropic behaviour of the white matter is investigated formance of the experiments, e.g., correct contact between upper
[21,32,79]. Consequently, samples of the CC and the CR were tested plate/plunger and tissue. To mimic physiological conditions and
as a function of axon alignment. prevent the samples from drying out, all experiments were per-
In a first step, the brains were carefully cut in the sagittal plane formed in transparent basins filled with 37 ± 2°C tempered DPBS
to separate the two hemispheres, see small subfigure in Fig. 1. solution. Due to the use of the solution, no change in sample shape
A total of one hundred and eighty one n = 181 cuboid samples due to gravity could be observed. Finally, to minimise the effects
were taken using a custom-made cutting tool, cf. Tyson et al. [80], of tissue degradation, all experiments were performed within 12 h
with an equidistant spacing of 4 mm and siliconised microtome post-mortem.
blades to prevent sample sticking. The corresponding sample di-
mensions can be found in the supplementary material. The num-
ber of specimen for the different deformation conditions (uniaxial 3.2.1. Axial tension experiments
tension/compression, semi-confined compression) and loading di- Fifty-seven specimens (n = 57, Table 1) were subjected to an
rections are listed in Table 1. orientation-dependent axial tensile test (index: at), see Fig. 2(a).
Specifically, two axon orientations (0°, 90°) were considered, with
the axons oriented parallel and perpendicular to the loading di-
3.2. Mechanical experiments rection. Samples were glued with superglue (Loctite® 406, Henkel
Ireland Operations and Research Ltd. Dublin, Ireland) to the up-
All experiments were conducted with two axial testing ma- per stamp and the lower plate, which were covered with sandpa-
chines (Zwick Z0.5, Zwick GmbH & Co. Ulm, Germany), cf. Fig. 2, per (grain size 320) for better adhesion. More precisely, we first
equipped with 2 N or 5 N load cells. All experiments were dis- glued one side of the sample to the sandpaper of the lower plate
placement controlled regardless of the deformation state and per- and moved the upper stamp down to glue the opposite side of
formed at a quasi-static deformation rate of 0.05 %/s. In addition, the sample to the stamp. After a curing period of 2–3 min, the
a camera was positioned in front of the specimens during the basin was filled with tempered DPBS solution. The stamp was then
test, providing images with a resolution of 2048 × 2048 pixels moved to the original height of the sample as measured before

3
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Fig. 2. Experimental setup for (a) axial tension, (b) axial compression, and (c) semi-confined compression experiments. Note, (i) and (ii) illustrate the undeformed and
deformed state, respectively. Scale bars: 10 mm.

gluing, and the mechanical test started. While the displacement u 3.3. Histological investigations
of the upper platen was predefined, the resulting force F was mea-
sured and converted to the mean engineering stress via P = F /Aat To correlate microstructural architecture and mechanical be-
by division through the axial cross-sectional area Aat , measured haviour, histological examinations were carried out on seventy-
from a digital image which was recorded before testing. The tensile four (nhist = 74) samples. We histologically examined the tissue ad-
stretch λat = 1 + u/h0 was calculated from u and the undeformed jacent to the mechanical sample. However, as the anatomical re-
sample height h0 . All experiments were performed up to λat = 1.5 gions are quite small, the adjacent tissue was limited and histo-
or until tissue failure occurring prior to this. logical samples could not be taken for each mechanical sample.
First, the preparations for the histological examinations were taken
3.2.2. Axial compression experiments from the immediate vicinity of the samples taken for the mechan-
Ninety-eight (n = 98) samples were subjected to orientation- ical tests. Samples were cryofixed by rapid freezing in isopentane
dependent (0°, 45°, 90°) axial compression (index: ac) tests. To di- (−160°C) cooled with liquid nitrogen (−196°C) and and thereafter
minish friction effects, polytetrafluoroethylene (PTFE) platens were stored in an ultra-low freezer at −80°C. For the preparation of the
glued on the upper custom-made light-weighted stamp and on actual histological sections (10 μm thickness), these were cut us-
the lower platen. The engineering stress P = F /Aac was determined ing a cryomicrotome at −15°C and stored in formal saline for up to
as ratio between the measured force F and the cross-sectional 3 h. To visualise myelin, glial cells, and neurons, the sections were
area Aac , measured from the undeformed sample. The compres- stained according to the Klüver-Barrera protocol, see Supplemen-
sive stretch λac = 1 − u/h0 was calculated from the displacement tary material, and preserved by a mounting medium and a cover
u of the upper stamp and the undeformed sample height h0 , cf. glass. Finally, all histological sections were digitised in two steps
Fig. 2(b). using a microscope (Nikon Eclipse Ni, Tokyo, Japan), equipped
with a 20× objective: First, overview images of the sample were
3.2.3. Semi-confined compression experiments scanned and second, areas of interest (ROI) of 0.65 × 0.65 mm
Based on the anisotropic results under compressive loading were randomly excised to avoid artifacts such as holes that would
shown in Section 3.2.2, we additionally performed twenty-six (n = distort the analysis.
26) semi-confined compression experiments (index: sc), sometimes
called ’plane strain compression’ [81] within the CC, see Fig. 2(c).
3.4. Data processing
The experimental protocol followed established protocols for mus-
cle and tendon tissue [82–84]. Mounting the brain specimens such
3.4.1. Analysis of the mechanical data
that the axons were parallel to the x-, y-, or z-direction, the axons
The measured force-displacement data were stored as raw data
were compressed (mode I), stretched (mode II), or kept constant
with a frequency of 10 Hz. Furthermore, the additional forces
in length (mode III), respectively. To account for variability in sam-
caused by the hydrostatic pressure in the DPBS solution were taken
ple dimensions, two identical test devices were used with 3 mm
into account by calculating the force-displacement data using a
spacing between the back and front plates.
reference curve representing the movement of the plunger in the
Remark 1 (Homogeneous deformations - effects of specimen ge- solution. To avoid local peaks that would potentially distort post-
ometry:). Generally, it is very difficult to realise ideal homoge- processing, the raw data were smoothed depending on the noise of
neous deformations on the very soft brain tissue, as boundary ef- the acquired data. Consequently, the smoothing (function smooth,
fects such as friction between the sample and the tool can influ- MATLAB® R2019b, The MathWorks, Inc.) was realised by calculat-
ence the deformation. Thus, slight deviations from the ideal homo- ing the mean value over 5%, 4%, and 4% strain for axial tension,
geneous state may occur. In order to approach these idealised con- compression, and semi-confined compression, respectively.
ditions, we have reduced the influence of the boundary conditions Pre-load used plays a special role, especially when sampling
by choosing a diameter-thickness ratio of ≤ 1 for the axial tension soft, biological tissue, since the curves are shifted both horizon-
experiments as proposed by Rashid et al. [42]. In case of the ax- tally and vertically. In previous experiments of brain tissue, differ-
ial compression and semi-confined compression experiments, we ent methods were used to account for pre-load: While most stud-
reduced the friction between the tools and between the tools and ies do not specify a specific pre-load [e.g. 12,15,22,45,86,87], other
the tissue by using materials with a smooth surface (polytetraflu- studies define the force origin as the point when the tool comes
oroethylene and polymethyl methacrylate) and by treating the sur- into contact with the tissue [e.g. 11,19,20,34,35,42,51,88]. Other au-
faces with silicone spray before inserting the tissue sample. In an thors use previously defined specific pre-load levels [37,46] or glue
early study [85], we showed that the friction between muscle tis- the sample to the tool [89]. Basically, it can be stated that all these
sue and tools can be significantly reduced in this way. methods have advantages and disadvantages and that a uniform

4
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Fig. 3. Clustering of (a) histological sections into (b) myelin, (c1 ) cells, and (d) unstained components. Further, (c2 ) illustrates the cell cluster with counted cells highlighted
by blue circles. Scale bars: 100 μm. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

and comparative sampling of brain tissue is only possible to a very peated three times to avoid local minima and the result with the
limited extent without the development of new methods. How- lowest variance was retained. The image was divided into 12 clus-
ever, since in the present study the experiments were performed ters using the steps described, which were manually assigned to
in DPBS solution, which caused additional noise, the use of a pre- microstructural components. Following Fig. 3, blue portions were
load is not advantageous. Therefore, the origin of the force/stress assigned to myelin, purple to cells, and white or yellow portions to
is defined as the point where the plunger comes into contact with unstained components. In (a), the ROI of a white matter sample is
the tissue sample. For this purpose, the point of contact was de- shown with the characteristic high proportion of myelin and glial
fined in a post-processing step based on the camera and force- cells aligned in the axon direction. In (b-d), the data points associ-
displacement data as the point at which the sample experiences ated with the corresponding cluster are highlighted in the original
a transverse strain (axial compression, semi-confined compression) colour, while the others are covered with a black mask. Thus, in
with a simultaneous significant change in the force signal. In the (b) only the blue elements are visible as the myelin portion, in (c1 )
tensile experiments, the origin of the force was defined by the ini- the cell cluster of the ROI is visible and in (d) the other uncoloured
tial sample height, which was measured optically before the test. components. Based on the clustered data points, the area fraction
Finally, the elastic regions of the force-displacement curves of the myelin is calculated.
were determined by successively computing tangents Ti by finite In addition, the cell clusters were used to determine the area
differences at the points ui , i = 1, 2, . . . , m, where m indicates the fraction ( f ) and density (ρ ) of cells, glial cells, and neurons per
maximum number of measuring points. The elastic range of the area. For this purpose, the clustered cell images were converted to
curve is defined as the region where the slope is monotonically greyscale in (c1 ) and binarised with rgb2gray or imbinarize, respec-
increasing, i.e., where Ti+1 ≥ Ti . The point u j at which T j+1 < T j tively. Further, clustering noise (small individual pixels that are not
marks the end point of the elastic region. identified as cells) was removed with bwareaopen (maximum pixel
In general, measurements were discarded from the analysis if size of 60) and truncated cells were deleted with imclearborder. Fi-
operational disturbances occurred, e.g., the sample slipping under nally, imfill and regionprops were used to fill the cells and evalu-
pressure or no force development despite contact and incomplete ate the regional properties, respectively. In (c2 ) the resulting bina-
adhesion of the sample under tensile load. Furthermore, we fol- rised image is shown. The counted cells are highlighted by blue cir-
lowed Tukey [90] and identified outliers based on the interquartile cles, with the corresponding equivalent diameter determined from
range. Considering the maximum stresses per deformation state, the regional properties. To distinguish between glia and neurons,
anatomical region, and axon orientation, we defined values that we introduced a threshold diameter dt . This simplified approach is
were more than 1.5 interquartile ranges above the 75% and below based on the mean cell density ρcells of the CX and the CC, repre-
the 25% quartile of the respective maximum stress as outliers. sentative of grey and white matter. We evaluated cell diameters for
each value between dt = 5–20 μm and calculated the cell density
by counting only cells whose equivalent diameter is smaller, equal
3.4.2. Processing of the histological data (ρcells≤dt = ρglial cells ), or larger (ρcells>dt = ρneurons ). Based on these
Using the histological sections described in Section 3.3, we de- two cell densities, we calculated the difference Qt between the cor-
termined the tissue components (myelin, cells) and the cell num- responding mean values of the CC and the CX for the tested diame-
bers (glia, neurons) in a post-processing step. To this end, the im- ters, hypothesising that the maximum difference marks the thresh-
ages of Klüver–Barrera stained sections were subjected to cluster- old diameter. As shown in Fig. 4, the maximum is at a diameter of
ing based on a k-means clustering algorithm of Lloyd [91] with 9.5 μm, indicating that the proportion of cells larger than 9.5 μm
seeds generated by the k-means++ algorithm of [92], which al- is significantly lower in the CC compared to the CX. Consequently,
lowed colour differentiation between myelin, cells, and unstained the cells larger than dt = 9.5 μm were counted as neurons and the
components. In more detail, the RGB images were converted to rest as glial cells. The corresponding area fraction was determined
the L × a × b colour space, resulting in a m × n× 3 matrix. By ex- by summing the counted cell areas and dividing by the ROI area.
cluding the lightness (L), the matrix reduces to m × n× 2 and the In a final step, we calculated the mean value of the area fractions
segmentation is based only on the colour spaces (a × b). This ma- (myelin, cells, glial cells, neurons) and densities (cells, glial cells,
trix of colour pixels was converted to greyscale and used as data neurons) over the defined ROIs.
points for k-means clustering, which involves the following steps:
Selecting the cluster number c, randomly selecting the first cluster
centre, selecting c − 1 cluster centres with D2 weighting, determin- 3.5. Statistics
ing the closest data points in the cluster using the usual Euclidean
norm, recalculating the centre based on the data points in the clus- For the experimental stress-strain curves, mean and standard
ter, and repeating the clustering until the changes in the centre deviation as well as coefficients of variance (CV) were determined
are less than the threshold of 10-6 [92–94]. This procedure was re- as the ratio of the standard deviation to the mean value depending

5
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

4.1. Mechanical material model for brain tissue

As described in Section 3.1, depending on the region, we char-


acterise brain tissue as either isotropic or transversely isotropic
material. Furthermore, the tissue exhibits strongly non-linear ma-
terial characteristics in combination with incompressible proper-
ties. Based on the well-known Ogden material model [96], Mero-
dio and Ogden [97] have developed a model that provides reliable
results [29] as well as a physical interpretation for the material pa-
rameters. The strain-energy function of this model
 +ψ
ψ =ψ iso aniso + U (1)
is decomposed into an isotropic, isochoric (ψ  ) and an anisotropic
iso
(ψaniso ) contribution. Further, the purely volumetric part
Kvol 2
U= (J − 1 − 2 ln(J )) (2)
4
Fig. 4. Cell density difference Qt for cells > dt (black line) and ≤ dt (green line) as a depends on J = det F , the determinant of the deformation gradient
function of the investigated border diameters dt (circles). (For interpretation of the F , and the parameter Kvol , controlling the degree of incompressibil-
references to colour in this figure legend, the reader is referred to the web version
ity. The isotropic, isochoric part
of this article.)

 = 2μ  α 
ψ λ1 + 
λα2 + 
λα3 − 3 (3)
iso
α2
on the anatomical regions and axon orientations. In addition, the
experimental data were statistically analysed for significant differ- is represented by the isotropic Ogden material model. Herein, λ ˆα =
i
ences between the maximum stresses of the investigated anatom- J −1 / 3 α
λi with i = 1, 2, 3 represent the isochoric parts of the princi-
ical regions and axon orientations. The analyses were performed pal stretches λα i
, the shear modulus is given by μ, and α defines a
for all three deformation modes by first performing a Shapiro- dimensionless constant. Finally, the anisotropic contribution of the
Wilk test for the maximum tensile, compressive and shear stresses, strain-energy function
followed by a Wilcoxon rank sum test. In addition, the Wilcoxon
2kμ  α /2 
rank sum test was used again to determine if there was a signif- ψaniso = I4 + 2I4−α /2 − 3 (4)
icant change in both microstructural properties (myelin, cell, glial α2

cell and neuron area fraction, and cell, glial cell and neuron den- depends on the anisotropy factor k and on the fourth invariant
sity) and material properties (μ, α ) for the different brain regions
I4 = tr[C M ]. (5)
and axon orientations. To quantify the relationship between mi-
crostructure and mechanics, the non-parametric Spearman rank Herein, C = FT F
defines the right Cauchy–Green tensor and M =
correlation coefficient rs was calculated. Thus, it can be determined m0  m0 is the second-order structural tensor, the dyadic product
to what extent a linear function can represent the correlation be- of the directional tensor m0 , representing the direction of the ax-
tween microstructure properties and material parameters. Accord- ons. In dependence on the various brain regions and their depen-
ing to the classification of Cohen [95], values |rs | ≥ 0.10 indicate a dence on axon orientation described in Section 5.2.2, we consider
weak, |rs | ≥ 0.30 a medium, and |rs | ≥ 0.50 a strong effect. For all the following strain-energy combinations
statistical tests, we calculated the p values and considered values 
ψ +ψ
aniso + U for CC
smaller than 0.05 to be significant. ψ = iso (6)
ψiso + U for CX, BG, CR.
4. Material modelling and parameter identification

4.2. Model parameter identification


Basically, the goal of a mathematical modelling approach is the
phenomenological description of the mechanical behaviour. In do-
To successfully identify the material parameters and thus cor-
ing so, the determined model parameters must ultimately apply
relate the mechanical properties with the microstructural compo-
to all deformation states. Here, the relationship between the mi-
nents, an inverse numerical optimisation method was used. Here,
crostructural information of the brain tissue and the mechanical
a forward finite element analysis is coupled with an optimisation
behaviour represented by the model parameters is investigated. For
algorithm that determines the optimal parameter vector p by min-
this purpose, a minimalist model, see also Remark 2, was used that
imising the error O ( p) between the measured and simulated re-
directly links the microstructural properties with the mechanical
sponses.
properties, but cannot necessarily describe all deformation states
For the special case of an isotropic material behaviour, the
with the same accuracy.
anisotropy factor k is zero, cf. Eq. (4). To identify the parameter
Remark 2 (Model selection:). Note, the choice of material model vector p = (μ, α , k = 0 ), the objective function
is limited to the objective of this study, which is to directly corre-
1  sim
m
late the mechanical properties with the microstructural properties. O CX,BG,CR ( p) = |Fj − Fjexp | (7)
Consequently, a model with a small number of material parame- m
j=1
ters representing the most important properties of the mechanical
behaviour, the strength, and the waveform was needed. Further- minimises the error between simulated (Fjsim ) and experimental
exp
more, the fit of the experimental data should be accurate to avoid (Fj ) data. This optimisation was carried out for each individual
errors due to an inappropriate approximation. Finally, we wanted measurement.
to use a model that could be compared with the current literature For the identification of the material parameters of the
to reveal possible similarities and discrepancies. anisotropic region CC, a two-step identification approach was

6
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

utilised: In the first step, the anisotropy factors were optimised us- However, the BG exhibits the highest variability, showing a maxi-
ing the following objective function mum of 89.7%. In contrast, the white matter regions (CR, CC) con-
sist mainly of myelinated axons, resulting in the highest propor-
O CC ( p) = O0CC◦ + O45
CC CC
◦ + O90◦ . (8) tions of myelin surface (mean: 78.1% in CC, 76.9% in CR). Statisti-
Herein, the sub-objective functions cal studies show that the CC and the CR contain significantly more
myelin than the BG and the CX.
1  sim
m
Following Section 3.4.2, we further differentiated between glial
OiCC ( p) = |Fj,i − F̄j,iexp | for i = 0◦ , 45◦ , 90◦ (9) cells and neurons using a threshold diameter (dt = 9.5 μm). In (b),
m
j=1
the distribution of the area fractions of glial cells and neurons
are functions of the orientation-dependent, mean experimental ( fglial cells , fneurons ) within the anatomical regions and the corre-
exp sim . This first identifica- sponding sum as cell area fraction ( fcells ) is shown. Within the lat-
curves F̄j,i and the simulated results Fj,i
tion step was performed for all three deformations modes, leading ter, region BG shows the highest variability with a minimum of
to kCC CC CC fcells = 2.4% and a maximum of fcells = 8.3%.
ac = 10.22, kat = 0.42, and ksc = 0.38. In a second step, these
parameters were fixed and the remaining parameters in p (μ, α ) Also, fcells decreases from the CX to the CR. While in the CX
were optimised via an average of 5.6% of the total area is covered by cells, in the CR
it is only 3.2%. Both white matter regions (CC, CR) differ signifi-
1  sim
m
cantly from the CX, and in addition the CR also from the BG. This
OiCC ( p) = |Fj,i − Fj,iexp | for i = 0◦ , 45◦ , 90◦ (10) decreasing trend of fcells can be attributed to the cell area cov-
m
j=1
ered by neurons, since fneurons is also highest in the CX and lowest
for each single measured force curve. The number of increments in the CR. Furthermore, the proportion of the neuron area of fcells
are m = 20, 100, and 100 for the tension, compression, and semi- is 70% in the CX and only 23% in the CR and 18% in the CC. The
confined compression experiments, respectively. For both compres- glial cells feature an opposite trend with the highest mean value
sion and shear tests, we assume homogeneous deformation. For of fglial cells = 3.1% in CC. The differences in the area fractions of
tension tests, this condition is not reasonable due to the glued sur- glial cells and neurons are statistically significant between white
faces. Therefore, we used one finite element for compression and matter (CC, CR) and the CX. Within fneurons the BG has a signif-
shear, and 250 elements for tension to discretise the samples. icantly higher area percentage than the CC and the CR. The glial
cells in turn cover a significantly larger area in the CC compared
to the BG and the latter compared to the CX.
5. Results The cell network densities ρcells , ρglialcells , and ρneurons are
shown in (c) as a function of the different regions. Similar to fcells ,
In the following, the mechanical behaviour within the investi- the region BG shows the highest variability, with a range from
gated anatomical regions is examined for the influence of the mi- ρcells = 688 to 1861 1/mm2 . In contrast, the CX has the lowest with
crostructure and thus the microstructural and mechanical proper- a mean of ρcells = 893 1/mm2 and the CC with ρcells = 1477 1/
ties are related to each other. mm2 features the highest number of cells per area. These region-
dependent differences are statistically significant. Specifically, the
5.1. Microstructural components cells density of the CX is significantly lower than the other three
regions, and the BG and CR are also significantly different from the
The studied anatomical regions (CX, BG, CC, CR) consist of dif- CC. Comparing the proportion of glial and neurons within ρcells , it
ferent microstructural components, cf. Section 2. Using the meth- becomes clear that small glial cells make up the largest proportion
ods described in Section 3.4.2, we quantified two main compo- in all anatomical regions. Thus, the number of glial cells in par-
nents in terms of area fraction f and density per area ρ , namely ticular influences the cell density. Consequently, the CC contains
myelin and cells (glia, neurons). While the former evaluates the the highest density of glial cells with a mean value of ρglial cells =
area covered by the components, the latter provides information 1415 1/mm2 . In addition, the statistical differences in glial cells be-
on the number of cells per area. Fig. 5(a) illustrates the myelin sur- tween the anatomical regions are identical to those in cell density.
face area fmyelin within the anatomical regions. As can be seen, the As a result, the neuron density in gray matter (CX) is significantly
gray matter tissue CX has no measurable myelin content. The min- higher than in white matter (CC, CR). In addition, the BG contains
imum of the mixed gray and white matter region BG is also 0%. significantly more neurons than the CC and the CR, with a 53.5%

Fig. 5. Region-dependent area fraction of (a) myelin, (b) cells, glial cells, and neurons as well as (c) density distribution of cells, glial cells, and neurons. Mean values are
shown as black crosses, medians as red straight lines, outliers as red plus, and the boxes as interquartile. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)

7
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Fig. 6. Orientation- and region-dependent stress-stretch behaviour from the tension experiments for the (a) BG, (b) CR0°, (c) CC0°, and (d) CC90°. Black curves including
circles indicate the mean experimental values, shaded areas the standard deviations, light gray the individual test curves, and the green curves indicate the fitted model
response. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

and 52.3% higher mean. Interestingly, the CX features a higher ab- are listed in Table 2. There are high variances (maximum CV of
solute density of glia (ρglial cells = 675 1/mm2 ) compared to neu- 50%) and similar stress magnitudes, with a difference between
rons (ρneurons = 218 1/mm2 ), although fglial cells is smaller than the highest (BG) and lowest voltage (CC90°) of only 36%. The lat-
fneurons . These results suggest that the white matter regions have a ter was statistically compared to CC0° at the comparative stretch
more densely packed cellular network than the gray matter. How- of 1.091 to investigate the degree of anisotropy. We observed no
ever, the area percentage of the cells in the gray matter is larger. significant difference and conclude that the tensile behaviour of
the CC is independent of axon orientation. Similarly, the stresses
5.2. Mechanical behaviour of the anatomical regions were compared by considering the CC
as isotropic to quantify the influence of the different microstruc-
5.2.1. Axial tension tural compositions, see Section 5.1. Despite the heterogeneous mi-
Fig. 6 shows the experimental stress-strain curves and model crostructure, were no significant differences between CC, BG, and
responses of the three regions (BG, CC, CR) as a function of axon CR, indicating a homogeneous material behaviour of the brain tis-
orientation. While the light grey curves represent the individual sue. Furthermore, we fitted the mechanical behaviour with the ma-
experiments with their individual maximum stretches, the black terial model presented in Section 3 and the mean curve of the
curves show the mean value of all experiments, the shaded areas simulated stress-strain response agrees well with the experimental
the standard deviations, and the green curve the model response. curves, see Fig. 6. Table 2 contains the mean material parameters
Both the experimental mean curves and the model responses are μat , αat , and kat as well as the coefficients of determination R2 .
plotted up to the minimum stretch level, i.e., the level reached Similar to the stress, the BG has the highest shear modulus with
by each sample. This level of strain, given by λat,max , is listed μat = 0.59 kPa. Note that the anisotropy factor kat of the CC is not
in Table 2. Regardless of region and axon orientation, all curves zero, although the stresses imply an isotropic behaviour. The sta-
are characterised by a pronounced nonlinear stress-strain response. tistical comparison of the material parameters (μat , αat ) shows no
The associated stresses for the comparable strain of λat = 1.071 significant differences.

Table 2
Maximum tensile stresses Pat,max at stretches λat,max , minimum stresses Pat,min at λat,min = 1.071, mean
material parameters (μat , αat , kat ) and coefficients of determination R2 in dependence on the region
(BG, CC, CR) and axon orientation.

Region Orient. [°] λat,max [−] Pat,λat =1.071 [kPa] μat [kPa] αat [−] kat [−] R2

BG – 1.073 0.11 ± 0.05 0.59 −8.6 0.00 0.925


CC 0 1.091 0.10 ± 0.05 0.34 −1.9 0.42 0.946
CC 90 1.106 0.07 ± 0.02 0.39 −7.6 0.42 0.915
CR 0 1.071 0.08 ± 0.03 0.42 −10.0 0.00 0.883

8
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Fig. 7. Orientation- and region-dependent stress-stretch behaviour from the compression experiments for the (a) CX, (b) BG, (c) CR, and the (d) CC. Solid curves including
circles indicate the mean experimental data, shaded areas the standard deviations, light coloured curves the individual tests, and the green curves indicate the fitted model
response. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

5.2.2. Axial compression entation, the Wilcoxon rank sum test was applied to statistically
Fig. 7 shows the stress-strain behaviour of the investigated compare the maximum stresses of 0°, 45°, and 90°. There is a
anatomical regions (CX, BG, CR, CC) as a function of axon orienta- trend in anisotropy within the CC region, as the differences be-
tion (0°, 45°, 90°) under compressive loading. All curves are shown tween all tested axon alignments of the CC, except between CC0°
up to a compression of λac,max = 0.55. Overall, the brain tissue and CC45°, are significantly different. When axons are loaded in
shows a non-linear, exponential behaviour with different character- tension (90°), the tissue responded 245% stiffer than in compres-
istics depending on region and axon orientation. The greatest vari- sion (0°) and 215% stiffer than in with axons under an angle of
ability is found at the point of maximum load. The corresponding 45°. The CR, in turn, exhibited a 15.4% stiffer response with axons
mean stresses and standard deviations are listed in Table 3. The oriented in the direction of loading, see (c). However, this differ-
anatomical region BG has the lowest coefficient of variation (CV) ence was not significant, indicating an isotropic material behaviour.
at 31.0%, while the region CC0° has the highest at 54.8%. CC90° Therefore, in the following, both axon orientations of the CR are
reaches the maximum mean stress of 1.45 kPa, and CR0° in turn combined and statistically compared to the other isotropic regions
the second highest value (1.05 kPa), see Fig. 7(d) and (c), respec- (CX, BG). The CR and BG are significantly stiffer than the CX with
tively. In contrast, CC45° and CC0° exhibit the softest response with 66.1% and 47.5% higher maximum stresses, respectively. These re-
maximum stresses of 0.46 and 0.42 kPa, see (d). To investigate sults suggest that the heterogeneous microstructure influences the
these stiffness differences concerning a dependence on axon ori- material behaviour of brain tissue under axial compression.
Table 3 summarises the mean material parameters and coef-
ficients of determination. The agreement between simulated and
Table 3
Maximum compression stresses Pac,max at λac,max = 0.55, mean material parameters
experimental stress-strain curves is satisfactory with a lowest R2
(μac , αac , kac ) and coefficients of determination R2 in dependence on the region (CX, of 0.994 and overlapping mean curves in Fig. 7. In terms of the
BG, CC, CR) and axon orientation. shear modulus μac , the mixed region BG of grey and white matter
Region Orient. [°] Pac,max [kPa] μac [kPa] αac [−] kac [−] R2
is the stiffest. Interestingly, the CC has the lowest shear moduli,
although the CC90° is characterised by the highest stresses. This
CX – 0.59 ± 0.19 0.11 8.7 0.00 0.997
is due to the high anisotropy factor kac , which reverses the stress
BG – 0.87 ± 0.27 0.16 8.9 0.00 0.994
CR 0 1.05 ± 0.44 0.14 10.2 0.00 0.996 trend. Conversely, the nonlinearity factor is highest in the white
CR 90 0.91 ± 0.35 0.14 10.2 0.00 0.996 matter regions (CC, CR) except in CC45°. Statistical comparisons of
CC 0 0.42 ± 0.23 0.05 11.2 10.22 0.995 the anisotropic (CC0,45,90°) and isotropic (CX, CR, BG) material pa-
CC 45 0.46 ± 0.24 0.09 8.4 10.22 0.994 rameters show significant differences for μac between CX and BG.
CC 90 1.45 ± 0.76 0.06 10.2 10.22 0.999
In addition, αac is significantly different for all orientations and re-

9
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Fig. 8. Semi-confined compression stress-stretch response as in dependence on axon direction represented by (a) mode I, (b) mode II, and (c) mode III. Black curves including
circles indicate the mean experimental values, shaded areas the standard deviations, light gray curves the individual tests, and the green curves indicate the fitted model
response. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

gions except for CX and BG, and for CC0° and CC90°. These re- 5.2.4. Comparison of deformation states with respect to region and
sults suggest that compressive stress waveforms depend on axonal axon orientation
alignment and the different microstructural compositions of gray Finally, the results of the different deformation states are com-
and white matter. pared with each other. For this purpose, Table 4 lists the mean
stresses and standard deviations at a comparable deformation
(minimum elastic stress strain) for all deformation states, regions,
5.2.3. Semi-confined compression and axon orientations. When comparing the CVs, it is noticeable
Fig. 8 shows the stress-strain relations of the semi-confined that the BG region has comparable values in tension and com-
compression experiments as a function of axon orientations. All ex- pression with 45.5% and 41.2%. However, the CC0° shows higher
periments were performed up to a maximum stretch of λsc,max = variability in compression (114.3%) and semi-confined compression
0.55 and the stress-stretch relations are characterised by a non- (75.0%) compared to tension (50.0%). CC90° and CR0° show a sim-
linear, exponential material behaviour. Axons are oriented in (a) ilar trend.
x-, (b) y-, and (c) z-directions, resulting in compression (mode I), Furthermore, there are differences in the stiffness of the tissue
extension (mode II), or constant length (mode III) of the ax- between the deformation states. Brain tissue shows a stiffer be-
ons under stress. The CV at the maximum load value is highest haviour when loaded under semi-confined compression than un-
for mode I at 51.9%, followed by 46.8% and 35.4% for mode III der compression. In the CC0°, the maximum compressive stress of
and mode II, respectively. The corresponding mean stresses are 0.007 kPa corresponds to only 17.5% of the stress in semi-confined
1.31 ± 0.68, 3.25 ± 1.15, and 3.44 ± 1.61 kPa for mode I, II and compression. Furthermore, the tensile stress is significantly higher
III, from lowest to highest. These differences are statistically sig- than the semi-confined compression and compression stresses. The
nificant, with modus I showing significantly softer response than latter behaviour is also referred to as TCA and indicates different
modi II and III, with the former showing a 148.1% stiffer response microstructural load transfer mechanisms in the two deformation
and the latter a 162.6% stiffer response. In contrast, no significant modes. To illustrate the TCA, Fig. 9 shows the ratio (Pat /|Pac |) be-
difference is found between modes II and III, suggesting that ad- tween the magnitudes of tensile stress Pat and compressive stress
ditional microstructural components besides axons contribute to |Pac | as a function of the absolute strain |u/h0 | for BG, CC0°, and
tissue stiffness. Finally, the responses of the model presented in CR0°. In general, the degree of TCA decreases with increasing de-
Section 4 can also be seen for all three modes. It can be seen that formation. This is particularly evident in the CC0° region. Smaller
the model response agrees well with the experimental response deformations show tensile stresses that are 180 times larger than
with a mean coefficient of determination of R2 = 0.992 ± 0.008. their compressive counterparts. This ratio decreases within 9.1%
Similar to the maximum stress, mode III has the highest mean strain to a TCA ratio of about 11. A similar behaviour is observed
shear modulus and mode I has the lowest value, with μsc = 0.31 for the CR0°, albeit less pronounced with a maximum TCA ratio of
and μsc = 0.27 kPa. In contrast, mode I has a mean nonlinearity 90 at smaller deformations. Finally, the BG also shows a decreas-
factor αsc of -4.41 and the other two modes 3.0 (mode II) and −1.8 ing TCA behaviour, albeit rather linear, with a maximum TCA ratio
(mode III). The statistical comparison of these parameters did not of 12 for smaller and 6 for larger deformations. In a further step,
reveal any significant differences. it was investigated whether the statistical differences discussed in

Table 4
Stresses of the three deformation states at comparable deformations (minimum elastic
tensile stretch) in dependence on region (CX, BG, CC, CR) and axon orientation.

Region Orientation [°] Pat,λat =1.071 [kPa] Pac,λac =0.929 [kPa] Psc,λsc =0.929 [kPa]

CX – – 0.015 ± 0.008 –
BG – 0.11 ± 0.05 0.017 ± 0.007 –
CR 0 0.08 ± 0.03 0.016 ± 0.007 –
CR 90 – 0.013 ± 0.004 –
CC 0 (modus I) 0.10 ± 0.05 0.007 ± 0.008 0.04 ± 0.03
CC 45 – 0.014 ± 0.008 –
CC 90 (modus II) 0.07 ± 0.02 0.016 ± 0.010 0.06 ± 0.02
CC 90 (modus III) – – 0.05 ± 0.03

10
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

sity (rs = −0.36). Furthermore, there is a mean effect between αac


and ρcells with rs = 0.41 ( p = 0.004). Interestingly, this correlations
arise from the number of glial cells, as both μac and αac correlate
significantly with ρglial cells (rs = −0.32 for μac , rs = 0.43 for αac )
and not significantly with ρneurons (rs = 0.15 for μac , rs = −0.23 for
αac ). Moreover, the area covered by glial cells correlates moder-
ately with the non-linearity factor αac , as highlighted in (f). Con-
versely, (e) shows the significant ( p = 0.03) negative correlation
between μac and fglial cells . The decrease in shear modulus with in-
creasing glial area and density is consistent with the negative cor-
relation with myelin area fraction, as glial cells form the myelin
sheath. Remarkably, the correlation of tensile stiffness with neu-
ronal area fraction shows a slightly non-significant (rs = 0.33 with
p = 0.09 for fneurons ), medium effect. These results imply that, al-
though compressive stiffness decreases with increasing glial cell
area, non-linearity increases. The latter also increases with increas-
Fig. 9. Development of tension-compression asymmetry. Tensile-to-compressive ing number of cells and glial cells. In contrast, tensile stiffness in-
stress ratio versus absolute strain for the BG, CC0°, and CR0°.
creases with increasing area covered by cells.

Sections 5.2.2 and 5.2.3 still hold for small deformations by sta- 6. Discussion
tistically comparing the stresses listed in Table 4. There were no
significant differences between the CX, BG, and CR under com- 6.1. Microstructural properties
pression, nor between the modes tested under semi-confined com-
pression. However, the CC90° remained significantly stiffer under In the present study, the area fractions of myelin and cells (glial
compression than the CC0° with a 128.6% higher stress. These re- cells, neurons) in the makeup of brain tissue microstructure were
sults indicate that, at small deformations, the influence of a hetero- quantified. We found that the area fraction of myelin is highest in
geneous microstructure on the mechanical behaviour can be ne- the white matter regions (CC, CR), as these are mainly composed
glected. However, for compressive loading, the axon orientation in of myelinated axons. This result is in line with studies of Budday
the CC needs to be considered, even for small deformations. et al. [72] and Antonovaite et al. [60]. Moreover, the mean values
of the CC and CR (78.5 and 79.6%) are in the range of 56–81% re-
5.3. Correlation of microstructure and mechanics ported for bovine white matter tissues [50]. Budday et al. [72], on
the other hand, have measured a value of only 15% in human tis-
To establish a direct correlation between the microstructural sue. The different processing of the histological data may cause this
architecture and the mechanical behaviour, the properties deter- discrepancy. In [72] a colour threshold was used, whereas in the
mined in the previous sections were correlated with each other. present study k-means clustering was used.
Specifically, Spearman’s rank correlation coefficients rs were cal- Strikingly, this study revealed an opposite tendency of the cell
culated between the mechanical properties in terms of material area percentage compared to cell density: On the one hand, the
parameters (μ, α ) and the microstructural components. A distinc- area covered with cells is largest in the grey matter and smallest
tion was made according to the deformation states uniaxial stress in the white matter. On the other hand, the cell density is lowest
(nhist,at = 28) and compression (nhist,ac = 46), as we assume dif- in the grey matter and highest in the white matter. These differ-
ferent load transfer mechanisms as evidenced by the observed ences result from the predominant cell type in each region, be-
TCA, see Section 5.2.4. For reasons of clarity, only the correlations cause most of the neurons that predominate in the grey matter
between the material parameters and the area fractions (myelin, are larger than the glial cells and occupy a larger area. However,
cells, glial cells, neurons) are highlighted in Fig. 10. The correla- the absolute number of glial cells is higher than the number of
tion of the material parameters with the densities (cells, glial cells, neurons. Therefore, the cell density is higher in the white matter,
neurons) and the table with Spearman’s rank correlation coeffi- where glial cells predominate. There are no studies that examine
cients rs can be found in the Supplementary material. Fig. 10(a) both the area fraction and the density of the cells. Begonia et al.
shows that the tensile and compressive shear modulus decreases [37] and Koser et al. [32] reported comparable area fractions in
with increasing myelin area fraction. The compressive correlation mixed pig tissue (7.85%) and mouse spinal cord (between 1.4% and
coefficient correspond to a moderate effect and is statistically sig- 12.7%), respectively. The former is within the range of 2.4 to 8.3%
nificant. In contrast, the nonlinearity factor αat correlates nega- we found for BG, and the latter is slightly higher than the white
tively with fmyelin (rs = −0.18), while αac has a significant posi- matter minimum (1.7%) and grey matter maximum (7.0%). Antono-
tive correlation (rs = 0.47 with p = 0.0 0 09), see (b). However, con- vaite et al. [60], in turn, determined values between 5 and 95%
sidering |αat |, tensile and compressive nonlinearity show a sim- for different subareas of the mouse hippocampus. The maximum
ilar trend. Regardless of the deformation state, the stiffness de- was measured in the densely packed granule cell layer, which is
creases and the nonlinearity increases as the area covered with not present in any of the anatomical regions we studied here. Bud-
myelin increases. Regarding the cell density and area fraction, a day et al. [72] also found that cell density was high in white mat-
positive correlation with the tensile shear modulus (rs = 0.36 for ter and low in grey matter. However, comparing the mean value
ρcells , rs = 0.40 for fcells ) can be observed. The latter correlation is of ρcells measured for BG, the present study shows a value 56.4%
significant at p = 0.04 and shown in (c). Although the maximum higher than documented in Budday et al. [72]. This difference is
stress in Section 5.2.1 has no regional dependencies, the stiffness due to the sampling location, as the BG includes several sub-areas
increases with the number of cells and the area covered by the with different proportions of grey matter. Budday et al. [72] chose
cells. In contrast, neither the tensile and compressive non-linearity the caudate nucleus, an area with a higher proportion of grey mat-
factor nor the compressive stiffness show significant correlations ter, than the putamen we chose for the present study. Therefore,
with cell area fraction, cf. (c) and (d). However, the compressive the number of neurons is higher and the overall cell density is
shear modulus significantly decreases with increasing cell den- lower.

11
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Fig. 10. Shear modules μat and μac and non-linearity factors αat and αac as a function of area fractions f of (a)-(b) myelin, (c)-(d) cells, (e)-(f) glial cells and (g)-(h)
neurons. Solid lines characterise linear regressions. The corresponding Spearman’s rank correlation coefficients are shown in each case in terms of uniaxial tension rs,at and
compression rs,ac . Note: ∗ marks p < 0.05.

12
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

Furthermore, we introduced a threshold diameter of dt = cept for the CX, which is not considered in the present study. In
9.5 μm to distinguish between neurons and glial cells. The sig- the other studies in Figure S1 (a) (see Supplementary material), the
nificant differences between the neuron density of the grey and brain tissue was loaded without defining an elastic range.
white matter underline the applicability of the chosen diameter, We also found a tension-compression asymmetry with up to
see Section 5.1. Since neurons are almost absent in the white mat- 180 times higher tensile stresses. However, other studies that ex-
ter, the significantly higher number of cells larger than dt in the CX amined tension and compression identified higher compressive
and BG indicates that the cells counted are neurons. To the best of stresses [23,26,27,33]. Two obvious reasons could lead to these
our knowledge, no study to date has reported an equivalent di- differences: Firstly, boundary conditions that lead to a deviation
ameter to distinguish between neurons and glial cells. We would from the assumed homogeneous deformation can cause discrepan-
like to point out that this approach cannot capture all sizes of glial cies in the results, see also Remark 1. Consequently, Rashid et al.
and neuronal cells present in the brain. This is partly due to the [42] have shown that the ratio of diameter to thickness of the
limited amount of histological data and partly due to the simpli- specimen tested in tension has a significant influence on the as-
fication of choosing one diameter for two main cell groups, which sumption of a homogeneous deformation state. The authors stated
themselves consist of several subtypes. Instead, it serves as an ef- that a diameter to thickness ratio ≤ 1 should be aimed for in or-
ficient first approximation to distinguish cell types based on their der to achieve a homogeneous stress distribution. Otherwise, the
size. influence of inhomogeneous deformation due to boundary condi-
tions such as bonding or clamping can no longer be neglected. In
6.2. Mechanical brain tissue properties this case, the measured forces, normalised to the initial surface, no
longer represent the stress ratios within the specimen. In the stud-
Generally, before the actual mechanical testing, tissues are sub- ies of Miller and Chinzei [27], Jin et al. [23], Budday et al. [26], and
jected to various mechanical and thermal procedures. While in Eskandari et al. [33], the TCA ratios are 3, 2.8, 1, and 1.3, respec-
some cases cadaver tissue is selected [16,18,26,51,86,98,99], in tively. In turn, we used a thickness/height ratio of 0.5. In addition,
other cases the tissue is taken as soon as possible after dissection Böl et al. [85] discusses the effect of friction between the speci-
[e.g. 12,36,50,59,79,100] in order to come as close as possible to men and the tool in uniaxial tests. The study reports an increase
the in vivo situation. Another important factor that significantly in stress response with increasing friction. Therefore, the degree
affects the mechanical response of the material is the loading of inhomogeneity is higher for glued samples as in Budday et al.
rate, which in the relevant literature varies between 0.0 0 064%/s [26] than in the current study with smooth contact surfaces that
[34] and 30 0,0 0 0%/s [36,39]. In addition, a variety of areas in the were additionally wetted with silicone oil. Another possible rea-
brain have been mechanically examined, with some studies sam- son is the loading rate. In the present study a rate of 0.05 %/s was
pling only a single region [11,13,41,54,79,98,101–103], while oth- used, which is lower than in the studies mentioned above. Assum-
ers examined up to seven regions [104]. Finally, the starting point ing incompressible material behaviour, the tissue exhibits higher
of the stress-strain relationship has a significant influence on the resistance at high strain rates than at low rates under compressive
maximum stress, but also on the elastic stretch range. Here, too, loading. This is due to poroelasticity, because the lower the strain
there are different approaches in the literature. Some studies do rate, the more fluid is pushed out of the tissue. Furthermore, the
not use a pre-load [e.g. 12,15,22,45,86,87, while others choose it studies of Miller and Chinzei [34] and Miller and Chinzei [27] have
arbitrarily [37,46] or select them visually depending on the defor- shown that the dependence on strain rate is more pronounced un-
mation [e.g., 11,19,20,34,35,42,51,88]. It is therefore not surprising der compression than under tension. Therefore, as shown in the
that the results obtained in these studies are highly dependent supplementary material, we measured lower compressive stresses
on the different protocols and vary accordingly. In Figure S1 (see than most studies in the literature.
Supplementary material), the maximum stresses reached at stretch Our study shows that the material behaviour can be consid-
of λ = 1.071 of the tensile (a) and at λ = 0.55 of the compressive ered homogeneous for small deformations, but becomes heteroge-
(b) stretch range from approximately 0.035 kPa [78] to 1.049 kPa neous as deformations increase. For large deformations, the CR of
[105] and −0.577 kPa [45] to −10.010 kPa [36], respectively. The the white matter behaves significantly stiffer than the CX and BG
mean tensile and compressive stresses of 0.091 ± 0.018 kPa and regions. These results confirm the findings of most previous stud-
−0.811 ± −0.202 kPa determined in the present study are thus ies that report higher stiffness of white matter compared to grey
within the range of values reported in the literature. The com- matter [16,23,29,36,43]. In contrast, Li et al. [44] found no signifi-
parison is based on the mean value of isotropic anatomical areas cant differences between the CX and the CR. A possible explanation
(CX, BG, CR), as a large proportion of studies have investigated for this could be that the regional difference is age-dependent, as
mixed tissue [e.g. 34,37,45,86] or white matter without consider- eight-week-old porcine brain tissue was used in the study. Another
ing different axon orientations [e.g. 20,28,35,44]. In the present study testing human brain tissue under compression and tension
study, the elastic stretch of brain tissue under tensile loading was showed an opposite trend with stiffer grey matter than white mat-
found to range from λ = 1.071 (CR) to λ = 1.106 (CC90°). This ter [26]. This inconsistency could also be due to age effects, as
range is consistent with the study by Bain and Meaney [106], re- samples with an average age of 66 years were used, and Elkin et al.
porting a conservative failure strain of λ = 1.14 for optical nerve [38] and MacManus et al. [52] reported a significant increase in
tissue. However, the elastic stretches observed in this study are stiffness with age.
lower than those of Miller and Chinzei [27] (λ = 1.21), Franceschini Interestingly, the present study showed anisotropic behaviour
et al. [28] (λ = 1.91 ± 0.42) and Velardi et al. [29] (λ = 1.27 for the of the CC for the deformation modes of uniaxial compression and
CR90° and λ = 1.63 for the CX). This discrepancy could be due to semi-confined compression, where axons loaded in tension dis-
the use of mixed white and grey matter, a different failure criterion played a higher stiffness than those loaded in compression. This
and tissue clamping instead of gluing. In addition, the CC with- finding is in contrast to previous studies that assumed that white
stands the highest tensile strain, with loading in the axon direction matter behaves isotropically under compressive loading [23,26,36].
leading to earlier failure than perpendicular to it, suggesting that However, these studies used tissues from 18-month-old cattle
axons fail earlier than the network of cells and the ECM in which to 70-year-old humans. Therefore, differences in species and age
they are embedded. These results mirror those of Velardi et al. could be the cause of the less pronounced anisotropy. In contrast,
[29] who also found that the CR0° fails at λ = 1.27,earlier than results on brain tissue under shear stress confirm the anisotropic
CR90° (λ = 1.46) and CC0°, resists the highest strain (λ = 1.52) ex- behaviour found in the present study [19,21,23].

13
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

It remains unclear whether axons are only loaded in tension, tent with the present study. In contrast, Antonovaite et al. [60] re-
because the semi-confined compression tests show a significant ported a negative correlation for both nuclei and neurons (Pear-
difference between mode I and III, where axons are loaded in com- son correlation coefficient of −0.73 and −0.81). To date, however,
pression or kept at a constant length. Therefore, the classical ap- no study has investigated the correlation for uniaxial tension. The
proach for transversely isotropic cardiovascular tissue [107], where pressure shear modulus, in turn, correlates negatively with the
fibres are only considered in tension, seems to be less appropri- area fraction and number of glial cells. Since glial cells form the
ate for nervous tissue. Unfortunately, to our knowledge, there is myelin around axons, this result corresponds to the negative cor-
no study that tests brain tissue under semi-confined compression relation with myelin area fraction. Conversely, Antonovaite et al.
conditions which we could use for comparison. There exists studies [60] showed an increase in memory modulus with increasing area
[82–84] that investigate this deformation mode on muscle and ten- covered by astrocytes, a specific type of glial cells (Pearson corre-
don tissue, but as the microstructure and stiffness of the brain dif- lation coefficient of 0.56). In this study, hippocampal white matter
fers significantly from these, a comparison is only of limited value. subregions were excluded from the correlation, while we exam-
ined the relationship for all anatomical regions. In addition, we ex-
6.3. Relationship between microstructural architecture and amined all glial cells without distinguishing between specific cell
mechanical properties types such as astrocytes.
Finally, the factor of compression non-linearity increases with
In the present study, we correlated microstructural architec- increasing myelin content. Importantly, the factor correlates signif-
ture and mechanical properties to explore the mechanisms of load icantly with cell density, glial cell density, and area fraction. Con-
transfer for different deformation modes. Other studies already sequently, white matter has higher non-linearity factors than grey
investigated the correlation between mechanics and microstruc- matter. This suggests that the compressive curve shape is directly
ture using cell area fraction [32,37,60,70], myelin area fraction influenced by the underlying myelin content and cell number. So
[50,60,72,77], cell density [72], neuron area fraction [60], and as- far, only Budday et al. [72] has investigated a comparable relation-
trocyte area fraction [60]. So far, however, there is no information ship and also found a positive correlation of the non-linearity fac-
on the influence of glial cells and neurons on the mechanical be- tor with cell density.
haviour at more than one deformation state, as Antonovaite et al.
[60] has only investigated the influence at indentations. Further- 6.4. Limitations of this study
more, correlation studies in the literature are mostly limited to one
or two anatomical regions, such as the white matter in Weicken- In this study, all tests were performed in vitro within 12 h
meier et al. [50,77], hippocampus and cortex in Antonovaite et al. post-mortem. While some studies [e.g. 45,49] report no effects
[60], or spinal cord in Koser et al. [32]. An exception is the study on mechanical properties during this post-mortem period, others
by Budday et al. [72], investigating CX, BG, CR, and CC, but based [e.g. 14,108,109] do. As in vivo testing methods such as magnetic
the correlation on the anatomical regions, only. To fill this gap, we resonance elastography are not able to characterise the full com-
investigated the direct influence of glial cells and neurons on me- plexity of the mechanical response of the brain, in vitro testing
chanical behaviour in different deformation states and independent currently remains a viable method to account for regional differ-
of the anatomical region. ences and behaviour under large deformations. However, in vitro
We found a negative correlation of compressive shear mod- studies like ours are limited to the staining protocol chosen by
ulus with myelin area fraction, suggesting that myelin does not Klüver Barrera. Since there is no staining method that makes all
contribute critically to the mechanical strength of the tissue. This components of the complex brain microstructure visible in one
result is consistent with a mouse brain study in which stiffness section, we decided to focus on the main components, cells and
measured by dynamic indentation was correlated with myelin myelin. Consequently, the chosen protocol makes cell bodies and
area fraction [60]. In contrast, the studies by Weickenmeier et al. myelin visible, while cell connections and synapses are not. Yet,
[50] and Budday et al. [72] document a positive correlation. How- image segmentation to identify cells and myelin is based on the
ever, both only looked at white matter tissue, whereas the present choice of the regions of interest. Here we used a large number of
study included all anatomical regions. Furthermore, the studies as- regions of interest to represent differences in microstructure. Due
sumed an isotropic material behaviour, whereas here, we observed to the limited anatomical size, it was not possible to histologi-
a pronounced anisotropy of the CC region. Therefore, our correla- cally examine every mechanical sample, so the number of samples
tion with myelin was negative, as the CC has a high proportion of was reduced from nmech = 181 to nhist = 74. Another limitation is
myelin, but low shear moduli compared to the isotropic anatomical the selection of the anisotropic Ogden model, a phenomenologi-
regions. cally motivated model that serves as a minimalist model to inves-
Interestingly, we found that the tensile shear modulus increases tigate the relationship between microstructural components and
with increasing area covered by cells, independent of cell type. mechanical behaviour. The experiments revealed different proper-
Since this correlation only exists under tensile stress, we suspect ties such as the significant anisotropy of CC under compressive
that the cell connections and not the cells themselves contribute loading, which is not significant under tensile loading, and the
to the mechanical response. Under tension, the cell connections to tensile-compressive asymmetry. These different properties lead to
neighbouring cells and nerve fibres can act as load-transfer chains, significantly different material parameters, such as the anisotropy
and the more chains are activated, the stiffer the response. The dis- factor k, which is higher under compression than under tension.
solution of these connections could explain the non-linear strain- Therefore, the deformation state was adjusted separately to deter-
stress response under tension, because after a high initial stiffness, mine a minimum error between experiment and simulation. De-
the curve flattens with increasing deformation. However, further spite the phenomenological nature of the present modelling ap-
studies of the inter-cellular network and its behaviour under ten- proach, the material parameter takes into account the mechanical
sile loading need to be carried out in the future to support this behaviour of the brain tissue. In this sense, μ stands for the gen-
assumption. Koser et al. [32], Thompson et al. [71], and Antono- eral material stiffness, α for the non-linearity of the stress-strain
vaite et al. [60] investigated the correlation between tissue stiff- response, and k for the degree of anisotropy due to structural ori-
ness measured by indentation and cell area fraction. Koser et al. entations. Although they are able to predict the stress-strain re-
[32] and Thompson et al. [71] found a positive correlation (both sponse under tension, some parameter configurations are physi-
with a Pearson correlation coefficient of 0.97), which is consis- cally problematic to interpret [110]. In particular, the stress re-

14
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

sponse in the corresponding strain energy function in Eq. (3) de- References
creases for negative values of α . In addition, the stress response
in tension results from the volume-preserving component in the [1] J. Bruns Jr, W.A. Hauser, The epidemiology of traumatic brain injury: a review,
Epilepsia 44 (s10) (2003) 2–10.
strain energy function in Eq. (1), as the material resistance in- [2] S. Mrozek, F. Vardon, T. Geeraerts, O. Bandschapp, Brain temperature: phys-
creases perpendicular to the direction of tension. Therefore, both iology and pathophysiology after brain injury, Anesthesiol. Res. Pract. 2012
the tensile and compressive behaviour is controlled by the com- (2012) 989487.
[3] F. Urbanek, M. Frink, O. Grün, R. Lohse, F. Hildebrand, Current opinions on
pressive resistance of the material. To overcome this problem, Mo- epidemiology, treatment and outcome after traumatic brain injury, J. Trauma
erman et al. [110] has presented an alternative approach with sep- Treat. 1 (2012). 2167–1222.
arate control of resistance to tension and compression, which could [4] K. Pogoda, L.K. Chin, P.C. Georges, F.R.J. Byfield, R. Bucki, R. Kim, M. Weaver,
R.G. Wells, C. Marcinkiewicz, P.A. Janmey, Compression stiffening of brain
be used in future studies. However, the present implementation of
and its effect on mechanosensing by glioma cells, New J. Phys. 16 (7) (2014)
the Ogden model is a commonly used formulation to represent 075002.
the mechanical properties of brain tissue. Therefore, choosing a [5] A.G. Kolias, P.J. Kirkpatrick, P.J. Hutchinson, Decompressive craniectomy: past,
present and future, Nat. Rev. Neurol. 9 (7) (2013) 405–415.
different model formulation would limit comparability with other
[6] J. Weickenmeier, C.A.M. Butler, P.G. Young, A. Goriely, E. Kuhl, The mechan-
studies and possible variations would remain undetected. Since es- ics of decompressive craniectomy: personalized simulations, Comput. Meth-
pecially in biomechanics with complex experimental setups, high ods Appl. Mech. Eng. 314 (2017) 180–195.
tissue variability, and complex microstructures, comparison with [7] M. Grujicic, G. Arakere, T. He, Material–modeling and structural–mechanics
aspects of the traumatic brain injury problem, Multidiscip. Model. Mater.
existing studies is a crucial issue, we accepted this limited inter- Struct. (2010).
pretability of α in tension. Furthermore, the experiments revealed [8] S. Chatelin, C. Deck, F. Renard, S. Kremer, C. Heinrich, J.P. Armspach, R. Will-
different properties such as the significant anisotropy of CC under inger, Computation of axonal elongation in head trauma finite element simu-
lation, J. Mech. Behav. Biomed. Mater. 4 (8) (2011) 1905–1919.
compressive loading, which is not significant under tensile loading, [9] R.J.H. Cloots, J.A.W. van Dommelen, M.G.D. Geers, A tissue-level anisotropic
and the TCA. These different properties lead to significantly differ- criterion for brain injury based on microstructural axonal deformation, J.
ent material parameters, such as the anisotropy factor k, which is Mech. Behav. Biomed. Mater. 5 (1) (2012) 41–52.
[10] S. Budday, C. Raybaud, E. Kuhl, A mechanical model predicts morphological
higher under compression than under tension. In order to obtain abnormalities in the developing human brain, Sci. Rep. 4 (5644) (2014) 1–7.
the most accurate model prediction for the stress-strain relation- [11] K.B. Arbogast, S.S. Margulies, Material characterization of the brainstem from
ship, we optimised a set of material parameters for each individ- oscillatory shear tests, J. Biomech. 31 (9) (1998) 801–807.
[12] K.L. Thibault, S.S. Margulies, Age-dependent material properties of the porcine
ual test, allowing us to investigate the influence of the microstruc-
cerebrum: effect on pediatric inertial head injury criteria, J. Biomech. 31 (12)
ture on the material response and its variability. Consequently, the (1998) 1119–1126.
determined material parameters are only able to represent the be- [13] L.E. Bilston, Z. Liu, N. Phan-Thien, Large strain behaviour of brain tissue in
shear: some experimental data and differential constitutive model, Biorheol-
haviour within the respective deformation state.
ogy 38 (4) (2001) 335–345.
[14] A. Garo, M. Hrapko, J.A.W. van Dommelen, G.W.M. Peters, Towards a reliable
7. Conclusion characterisation of the mechanical behaviour of brain tissue: the effects of
post-mortem time and sample preparation, Biorheology 44 (1) (2007) 51–
58.
The present study shows that the strongly non-linear mechan- [15] S. Chatelin, J. Vappou, S. Roth, J.S. Raul, R. Willinger, Towards child versus
ical behaviour of nervous tissue is a direct result of the hetero- adult brain mechanical properties, J. Mech. Behav. Biomed. Mater. 6 (2012)
geneous tissue microstructure. Although white matter tissue con- 166–173.
[16] A.E. Forte, S.M. Gentleman, D. Dini, On the characterization of the het-
tains more myelin and cells than grey matter, its shear moduli are erogeneous mechanical response of human brain tissue, Biomech. Model.
smaller. This suggests that there is a negative correlation between Mechanobiol. 16 (3) (2017) 907–920.
the area fraction of myelin and glial cells and the shear stiffness. [17] A. Tabet, S. Mommer, J.A. Vigil, C. Hallou, H. Bulstrode, O.A. Scherman, Me-
chanical characterization of human brain tissue and soft dynamic gels ex-
In contrast, the area fraction of cells, neurons and glial cells cor- hibiting electromechanical neuro-mimicry, Adv. Healthc. Mater. 8 (10) (2019)
relates positively with the tensile shear stiffness, suggesting that 190 0 068.
the number of cell connections is responsible for the tissue stiff- [18] K.K. Darvish, J.R. Crandall, Nonlinear viscoelastic effects in oscillatory shear
deformation of brain tissue, Med. Eng. Phys. 23 (9) (2001) 633–645.
ness. Our study also found that myelin and glial cell area and cell [19] M.T. Prange, S.S. Margulies, Regional, directional, and age-dependent prop-
density influence the non-linearity of the stress-strain response, as erties of the brain undergoing large deformation, J. Biomech. Eng. 124 (2)
the non-linearity of white matter differs significantly from that of (2002) 244–252.
[20] M. Hrapko, J.A.W. van Dommelen, G.W.M. Peters, J.S.H.M. Wismans, Character-
grey matter. Understanding the microstructural load transfer mech-
isation of the mechanical behaviour of brain tissue in compression and shear,
anisms in the brain is a crucial step towards developing reliable Biorheology 45 (6) (2008) 663–676.
computational models to simulate surgical procedures, e.g., cancer [21] Y. Feng, R.J. Okamoto, R. Namani, G.M. Genin, P.V. Bayly, Measurements
tissue removal. To make brain models more accurate in the future, of mechanical anisotropy in brain tissue and implications for transversely
isotropic material models of white matter, J. Mech. Behav. Biomed. Mater. 23
our study suggests that it is critical to include microstructural in- (2013) 117–132.
formation into the material characterisation of nervous tissue. [22] M.D. Gilchrist, B. Rashid, J.G. Murphy, G. Saccomandi, Quasi-static deforma-
tions of biological soft tissue, Math. Mech. Solids 18 (6) (2013) 622–633.
[23] X. Jin, F. Zhu, H. Mao, M. Shen, K.H. Yang, A comprehensive experimental
Declaration of Competing Interest study on material properties of human brain tissue, J. Biomech. 46 (16) (2013)
2795–2801.
The authors declare that they have no known competing finan- [24] B. Rashid, M. Destrade, M.D. Gilchrist, Mechanical characterization of brain
tissue in simple shear at dynamic strain rates, J. Mech. Behav. Biomed. Mater.
cial interests or personal relationships that could have appeared to 28 (2013) 71–85.
influence the work reported in this paper. [25] M. Destrade, M.D. Gilchrist, J.G. Murphy, B. Rashid, G. Saccomandi, Extreme
softness of brain matter in simple shear, Int. J. Non Linear Mech. 75 (2015)
54–58.
Acknowledgements [26] S. Budday, G. Sommer, C. Birkl, C. Langkammer, J. Haybaeck, J. Kohnert,
M. Bauer, F. Paulsen, P. Steinmann, E. Kuhl, G.A. Holzapfel, Mechanical char-
Partial support for this research was provided by the Deutsche acterization of human brain tissue, Acta Biomater. 48 (2017) 319–340.
[27] K. Miller, K. Chinzei, Mechanical properties of brain tissue in tension, J.
Forschungsgemeinschaft (DFG) under Grants no. 404568779.
Biomech. 35 (4) (2002) 483–490.
[28] G. Franceschini, D. Bigoni, P. Regitnig, G.A. Holzapfel, Brain tissue deforms
Supplementary material similarly to filled elastomers and follows consolidation theory, J. Mech. Phys.
Solids 54 (12) (2006) 2592–2620.
[29] F. Velardi, F. Fraternali, M. Angelillo, Anisotropic constitutive equations and
Supplementary material associated with this article can be experimental tensile behavior of brain tissue, Biomech. Model. Mechanobiol.
found, in the online version, at doi:10.1016/j.actbio.2022.08.034. 5 (1) (2006) 53–61.

15
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

[30] A. Karimi, M. Navidbakhsh, H. Yousefi, A. Montevalli Haghi, S.J.A. Sadati, RE- [61] A. Goriely, M.G.D. Geers, G.A. Holzapfel, J. Jayamohan, A. Jérusalem, S. Sival-
TRACTED: Experimental and numerical study on the mechanical behavior of oganathan, W. Squier, J.A.W. van Dommelen, S. Waters, E. Kuhl, Mechanics
rat brain tissue, Perfusion 29 (4) (2014) 307–314. of the brain: perspectives, challenges, and opportunities, Biomech. Model.
[31] B. Rashid, M. Destrade, M.D. Gilchrist, Mechanical characterization of brain Mechanobiol. 14 (5) (2015) 931–965.
tissue in tension at dynamic strain rates, J. Mech. Behav. Biomed. Mater. 33 [62] R. Wang, M. Sarntinoranont, Biphasic analysis of rat brain slices under creep
(2014) 43–54. indentation shows nonlinear tension-compression behavior, J. Mech. Behav.
[32] D.E. Koser, E. Moeendarbary, J. Hanne, S. Kuerten, K. Franze, CNS cell distribu- Biomed. Mater. 89 (2019) 1–8.
tion and axon orientation determine local spinal cord mechanical properties, [63] N. Abolfathi, A. Naik, M.S. Chafi, G. Karami, M. Ziejewski, A micromechani-
Biophys. J. 108 (9) (2015) 2137–2147. cal procedure for modelling the anisotropic mechanical properties of brain
[33] F. Eskandari, M. Shafieian, M.M. Aghdam, K. Laksari, Tension strain-soften- white matter, Comput. Methods Biomech. Biomed. Eng. 12 (3) (2009) 249–
ing and compression strain-stiffening behavior of brain white matter, Ann. 262.
Biomed. Eng. 49 (1) (2021) 276–286. [64] G. Karami, N. Grundman, N. Abolfathi, A. Naik, M. Ziejewski, A micromechan-
[34] K. Miller, K. Chinzei, Constitutive modelling of brain tissue: experiment and ical hyperelastic modeling of brain white matter under large deformation, J.
theory, J. Biomech. 30 (11) (1997) 1115–1121. Mech. Behav. Biomed. Mater. 2 (3) (2009) 243–254.
[35] S. Cheng, L.E. Bilston, Unconfined compression of white matter, J. Biomech. [65] K.B. Bernick, T.P. Prevost, S. Suresh, S. Socrate, Biomechanics of single cortical
40 (1) (2007) 117–124. neurons, Acta Biomater. 7 (3) (2011) 1210–1219.
[36] F. Pervin, W.W. Chen, Dynamic mechanical response of bovine gray matter [66] S. Javid, A. Rezaei, G. Karami, A micromechanical procedure for viscoelas-
and white matter brain tissues under compression, J. Biomech. 42 (6) (2009) tic characterization of the axons and ECM of the brainstem, J. Mech. Behav.
731–735. Biomed. Mater. 30 (2014) 290–299.
[37] M.T. Begonia, R. Prabhu, J. Liao, M.F. Horstemeyer, L.N. Williams, The influence [67] L. Bollmann, D.E. Koser, R. Shahapure, H.O.B. Gautier, G.A. Holzapfel, G. Scar-
of strain rate dependency on the structure - property relations of porcine celli, M.C. Gather, E. Ulbricht, K. Franze, Microglia mechanics: immune activa-
brain, Ann. Biomed. Eng. 38 (10) (2010) 3043–3057. tion alters traction forces and durotaxis, Front. Cell. Neurosci. 9 (2015) 1–16
[38] B.S. Elkin, A. Ilankovan, B.I. Morrison, Age-dependent regional mechanical 363.
properties of the rat hippocampus and cortex, J. Biomech. Eng. 132 (1) (2010) [68] R. de Rooij, K.E. Miller, E. Kuhl, Modeling molecular mechanisms in the axon,
1–10. Comput. Mech. 59 (3) (2017) 523–537.
[39] F. Pervin, W.W. Chen, Effect of inter-species, gender, and breeding on the me- [69] M. Anthonisen, M. Rigby, M.H. Sangji, X.Y. Chua, P. Grütter, Response of me-
chanical behavior of brain tissue, NeuroImage 54 (2011) S98–S102. chanically-created neurites to extension, J. Mech. Behav. Biomed. Mater. 98
[40] T.P. Prevost, G. Jin, M.A. de Moya, H.B. Alam, S. Suresh, S. Socrate, Dynamic (2019) 121–130.
mechanical response of brain tissue in indentation in vivo, in situ and in [70] N. Antonovaite, S.V. Beekmans, E.M. Hol, W.J. Wadman, D. Iannuzzi,
vitro, Acta Biomater. 7 (12) (2011) 4090–4101. Structure-stiffness relation of live mouse brain tissue determined by
[41] K. Laksari, M. Shafieian, K. Darvish, Constitutive model for brain tissue under depth-controlled indentation mapping, arXiv preprint arXiv:1802.02245
finite compression, J. Biomech. 45 (4) (2012) 642–646. (2018).
[42] B. Rashid, M. Destrade, M.D. Gilchrist, Mechanical characterization of brain [71] A.J. Thompson, E.K. Pillai, I.B. Dimov, S.K. Foster, C.E. Holt, K. Franze, C.A. Ma-
tissue in compression at dynamic strain rates, J. Mech. Behav. Biomed. Mater. son, M.E. Bronner, P. Bovolenta, V. Castellani, Rapid changes in tissue mechan-
10 (2012) 23–38. ics regulate cell behaviour in the developing embryonic brain, eLife 8 (2019)
[43] W. Zhang, L. Liu, Y. Xiong, Y. Liu, S. Yu, C. Wu, W. Guo, Effect of in vitro E39356.
storage duration on measured mechanical properties of brain tissue, Sci. Rep. [72] S. Budday, M. Sarem, L. Starck, G. Sommer, J. Pfefferle, N. Phunchago, E. Kuhl,
8 (1) (2018) 1247. F. Paulsen, P. Steinmann, V.P. Shastri, G.A. Holzapfel, Towards microstruc-
[44] Z. Li, H. Yang, G. Wang, X. Han, S. Zhang, Compressive properties and con- ture-informed material models for human brain tissue, Acta Biomater. 104
stitutive modeling of different regions of 8-week-old pediatric porcine brain (2020) 53–65.
under large strain and wide strain rates, J. Mech. Behav. Biomed. Mater. 89 [73] S. Herculano-Houzel, The human brain in numbers: a linearly scaled-up pri-
(2019) 122–131. mate brain, Front. Hum. Neurosci. 3 (2009) 31.
[45] D. Singh, S. Boakye-Yiadom, D.S. Cronin, Comparison of porcine brain me- [74] R. Lüllmann-Rauch, E. Asan, Taschenlehrbuch Histologie, Thieme, Stuttgart,
chanical properties to potential tissue simulant materials in quasi-static and 2019.
sinusoidal compression, J. Biomech. 92 (2019) 84–91. [75] L.W. Lau, R. Cua, M.B. Keough, S. Haylock-Jacobs, V.W. Yong, Pathophysiology
[46] T.P. Prevost, A. Balakrishnan, S. Suresh, S. Socrate, Biomechanics of brain tis- of the brain extracellular matrix: a new target for remyelination, Nat. Rev.
sue, Acta Biomater. 7 (1) (2011) 83–95. Neurosci. 14 (10) (2013) 722–729.
[47] J.A.W. van Dommelen, T.P.J. van der Sande, M. Hrapko, G.W.M. Peters, Me- [76] P. Sauleau, E. Lapouble, D. Val-Laillet, C.H. Malbert, The pig model in brain
chanical properties of brain tissue by indentation: interregional variation, J. imaging and neurosurgery, Animal 3 (8) (2009) 1138–1151.
Mech. Behav. Biomed. Mater. 3 (2) (2010) 158–166. [77] J. Weickenmeier, M. Kurt, E. Ozkaya, M. Wintermark, K.B. Pauly, E. Kuhl, Mag-
[48] S.J. Lee, M.A. King, J. Sun, H.K. Xie, G. Subhash, M. Sarntinoranont, Measure- netic resonance elastography of the brain: acomparison between pigs and hu-
ment of viscoelastic properties in multiple anatomical regions of acute rat mans, J. Mech. Behav. Biomed. Mater. 77 (2018) 702–710.
brain tissue slices, J. Mech. Behav. Biomed. Mater. 29 (2014) 213–224. [78] F. Eskandari, M. Shafieian, M.M. Aghdam, K. Laksari, Structural anisotropy
[49] S. Budday, R. Nay, R. de Rooij, P. Steinmann, T. Wyrobek, T.C. Ovaert, E. Kuhl, vs. mechanical anisotropy: the contribution of axonal fibers to the mate-
Mechanical properties of gray and white matter brain tissue by indentation, rial properties of brain white matter, Ann. Biomed. Eng. 49 (3) (2021) 991–
J. Mech. Behav. Biomed. Mater. 46 (2015) 318–330. 999.
[50] J. Weickenmeier, R. de Rooij, S. Budday, P. Steinmann, T.C. Ovaert, E. Kuhl, [79] Y. Feng, C.-H. Lee, L. Sun, S. Ji, X. Zhao, Characterizing white matter tissue in
Brain stiffness increases with myelin content, Acta Biomater. 42 (2016) large strain via asymmetric indentation and inverse finite element modeling,
265–272. J. Mech. Behav. Biomed. Mater. 65 (2017) 490–501.
[51] A. Samadi-Dooki, G.Z. Voyiadjis, R.W. Stout, An indirect indentation method [80] A.L. Tyson, S.T. Hilton, L.C. Andreae, Rapid, simple and inexpensive production
for evaluating the linear viscoelastic properties of the brain tissue, J. Biomech. of custom 3Dprinted equipment for large-volume fluorescence microscopy,
Eng. 139 (6) (2017) 1–12. Int. J. Pharm. 494 (2) (2015) 651–656.
[52] D.B. MacManus, J.G. Murphy, M.D. Gilchrist, Mechanical characterisation of [81] E.M. Arruda, M.C. Boyce, A three-dimensional constitutive model for the large
brain tissue up to 35% strain at 1, 10, and 100/s using a custom-built mi- stretch behavior of rubber elastic materials, J. Mech. Phys. Solids 41 (2) (1993)
cro-indentation apparatus, J. Mech. Behav. Biomed. Mater. 87 (2018) 256–266. 389–412.
[53] A.S. Mijailovic, B. Qing, D. Fortunato, K.J. van Vliet, Characterizing viscoelas- [82] M. Böl, A.E. Ehret, K. Leichsenring, C. Weichert, R. Kruse, On the anisotropy
tic mechanical properties of highly compliant polymers and biological tissues of skeletal muscle tissue under compression, Acta Biomater. 10 (7) (2014)
using impact indentation, Acta Biomater. 71 (2018) 388–397. 3225–3234.
[54] L. Qian, H. Zhao, Y. Guo, Y. Li, M. Zhou, L. Yang, Z. Wang, Y. Sun, Influence of [83] M. Böl, A.E. Ehret, K. Leichsenring, M. Ernst, Tissue-scale anisotropy and com-
strain rate on indentation response of porcine brain, J. Mech. Behav. Biomed. pressibility of tendon in semi-confined compression tests, J. Biomech. 48 (6)
Mater. 82 (2018) 210–217. (2015) 1092–1098.
[55] L.Z. Shuck, S.H. Advani, Rheological response of human brain tissue in, J. Basic [84] M. Böl, K. Leichsenring, M. Ernst, A.E. Ehret, Long-term mechanical behaviour
Eng. 94 (4) (1972) 905–911. of skeletal muscle tissue in semi-confined compression experiments, J. Mech.
[56] J.E. Galford, J.H. McElhaney, A viscoelastic study of scalp, brain, and dura, J. Behav. Biomed. Mater. 63 (2016) 115–124.
Biomech. 3 (2) (1970) 211–221. [85] M. Böl, R. Kruse, A.E. Ehret, K. Leichsenring, T. Siebert, Compressive proper-
[57] H. Metz, J. McElhaney, A.K. Ommaya, A comparison of the elasticity of live, ties of passive skeletal muscle—the impact of precise sample geometry on
dead, and fixed brain tissue, J. Biomech. 3 (4) (1970) 453–458. parameter identification in inverse finite element analysis, J. Biomech. 45 (15)
[58] A.V. Shulyakov, S.S. Cenkowski, R.J. Buist, M.R.D. Bigio, Age-dependence of in- (2012) 2673–2679.
tracranial viscoelastic properties in living rats, J. Mech. Behav. Biomed. Mater. [86] R.K. Prabhu, M.T. Begonia, W.R. Whittington, M.A. Murphy, Y. Mao, J. Liao,
4 (3) (2011) 484–497. L.N. Williams, M.F. Horstemeyer, J. Sheng, Compressive mechanical properties
[59] J.D. Finan, B.S. Elkin, E.M. Pearson, I.L. Kalbian, B. Morrison, Viscoelastic prop- of porcine brain: experimentation and modeling of the tissue hydration ef-
erties of the rat brain in the sagittal plane: effects of anatomical structure fects, Bioengineering 6 (2) (2019) 40.
and age, Ann. Biomed. Eng. 40 (1) (2012) 70–78. [87] N. Reiter, B. Roy, F. Paulsen, S. Budday, Insights into the microstructural origin
[60] N. Antonovaite, L.A. Hulshof, E.M. Hol, W.J. Wadman, D. Iannuzzi, Viscoelastic of brain viscoelasticity, J. Elast. 145 (1) (2021) 99–116.
mapping of mouse brain tissue: relation to structure and age, J. Mech. Behav. [88] A. Gefen, S.S. Margulies, Are in vivo and in situ brain tissues mechanically
Biomed. Mater. 113 (2021) 104159. similar? J. Biomech. 37 (9) (2004) 1339–1352.

16
JID: ACTBIO
ARTICLE IN PRESS [m5G;August 26, 2022;19:6]

M. Hoppstädter, D. Püllmann, R. Seydewitz et al. Acta Biomaterialia xxx (xxxx) xxx

[89] S. Budday, G. Sommer, J. Haybaeck, P. Steinmann, G.A. Holzapfel, E. Kuhl, [101] B. Coats, S.S. Margulies, Material properties of porcine parietal cortex, J.
Rheological characterization of human brain tissue, Acta Biomater. 60 (2017) Biomech. 39 (13) (2006) 2521–2525.
315–329. [102] A. Gefen, N. Gefen, Q. Zhu, R. Raghupathi, S.S. Margulies, Age-dependent
[90] J.W. Tukey, Exploratory data Analysis, Addison-Wesley Series in Behavioral changes in material properties of the brain and braincase of the rat, J. Neuro-
Sciences vol. 688 (1977) 581–582. trauma 20 (11) (2003) 1163–1177.
[91] S. Lloyd, Least squares quantization in PCM, IEEE Trans. Inf. Theory 28 (2) [103] K.M. Labus, C.M. Puttlitz, Viscoelasticity of brain corpus callosum in biaxial
(1982) 129–137. tension, J. Mech. Phys. Solids 96 (2016) 591–604.
[92] D. Arthur, S. Vassilvitskii, K-means++: the advantages of careful seeding, in: [104] D.B. MacManus, B. Pierrat, J.G. Murphy, M.D. Gilchrist, Region and species de-
Proceedings of the Eighteenth Annual ACM-SIAM Symposium on Discrete Al- pendent mechanical properties of adolescent and young adult brain tissue,
gorithms, 2007, pp. 1027–1035. Sci. Rep. 7 (1) (2017) 13729.
[93] D. Pollard, Strong consistency of k-means clustering, Ann. Stat. 9 (1) (1981) [105] B. Rashid, M. Destrade, M.D. Gilchrist, Inhomogeneous deformation of brain
135–140. tissue during tension tests, Comput. Mater. Sci. 64 (2012) 295–300.
[94] T. Kanungo, D.M. Mount, N.S. Netanyahu, C.D. Piatko, R. Silverman, A.Y. Wu, [106] A.C. Bain, D.F. Meaney, Tissue-level thresholds for axonal damage in an ex-
An efficient k-means clustering algorithm: analysis and implementation, IEEE perimental model of central nervous system white matter injury, J. Biomech.
Trans. Pattern Anal. Mach. Intell. 24 (7) (2002) 881–892. Eng. 122 (6) (20 0 0) 615–622.
[95] J. Cohen, Statistical Power Analysis for the Behavioral Sciences, Academic [107] T.C. Gasser, R.W. Ogden, G.A. Holzapfel, Hyperelastic modelling of arterial
Press, 2013. layers with distributed collagen fibre orientations, J. R. Soc. Interface 3 (6)
[96] R.W. Ogden, Large deformation isotropic elasticity – on the correlation of the- (2006) 15–35.
ory and experiment for incompressible rubberlike solids, Proc. R. Soc. Lond. A [108] J. Vappou, E. Breton, P. Choquet, R. Willinger, A. Constantinesco, Assessment
326 (1567) (1972) 565–584. of in vivo and post-mortem mechanical behavior of brain tissue using mag-
[97] J. Merodio, R.W. Ogden, Mechanical response of fiber-reinforced incompress- netic resonance elastography, J. Biomech. 41 (14) (2008) 2954–2959.
ible non-linearly elastic solids, Int. J. Non Linear Mech. 40 (2) (2005) 213–227. [109] J. Weickenmeier, M. Kurt, E. Ozkaya, R. de Rooij, T.C. Ovaert, R.L. Ehman,
[98] S. Nicolle, M. Lounis, R. Willinger, J.F. Palierne, Shear linear behavior of brain K.B. Pauly, E. Kuhl, Brain stiffens post mortem, J. Mech. Behav. Biomed. Mater.
tissue over a large frequency range, Biorheology 42 (3) (2005) 209–223. 84 (2018) 88–98.
[99] F. Shen, T.E. Tay, J.Z. Li, S. Nigen, P.V.S. Lee, H.K. Chan, Modified bilston non- [110] K.M. Moerman, C.K. Simms, T. Nagel, Control of tension–compression asym-
linear viscoelastic model for finite element head injury studies, J. Biomech. metry in Ogden hyperelasticity with application to soft tissue modelling, J.
Eng. 128 (5) (2006) 797–801. Mech. Behav. Biomed. Mater. 56 (2016) 218–228.
[100] B.S. Elkin, A. Ilankova, B. Morrison, Dynamic, regional mechanical properties
of the porcine brain: indentation in the coronal plane, J. Biomech. Eng. 133
(7) (2011) 1–7.

17

You might also like