ArticlePDF Available

Theoretical and experimental investigation of near-infrared light propagation in a model of the adult head

Optica Publishing Group
Applied Optics
Authors:

Abstract and Figures

Near-infrared light propagation in various models of the adult head is analyzed by both time-of-flight measurements and mathematical prediction. The models consist of three- or four-layered slabs, the latter incorporating a clear cerebrospinal fluid (CSF) layer. The most sophisticated model also incorporates slots that imitate sulci on the brain surface. For each model, the experimentally measured mean optical path length as a function of source-detector spacing agrees well with predictions from either a Monte Carlo model or a finite-element method based on diffusion theory or a hybrid radiosity-diffusion theory. Light propagation in the adult head is shown to be highly affected by the presence of the clear CSF layer, and both the optical path length and the spatial sensitivity profile of the models with a CSF layer are quite different from those without the CSF layer. However, the geometry of the sulci and the boundary between the gray and the white matter have little effect on the detected light distribution.
Content may be subject to copyright.
Theoretical and experimental investigation of
near-infrared light propagation in a model of the
adult head
Eiji Okada, Michael Firbank, Martin Schweiger, Simon R. Arridge, Mark Cope, and
David T. Delpy
Near-infrared light propagation in various models of the adult head is analyzed by both time-of-flight
measurements and mathematical prediction. The models consist of three- or four-layered slabs, the
latter incorporating a clear cerebrospinal fluid ~CSF! layer. The most sophisticated model also incor-
porates slots that imitate sulci on the brain surface. For each model, the experimentally measured mean
optical path length as a function of source–detector spacing agrees well with predictions from either a
Monte Carlo model or a finite-element method based on diffusion theory or a hybrid radiosity–diffusion
theory. Light propagation in the adult head is shown to be highly affected by the presence of the clear
CSF layer, and both the optical path length and the spatial sensitivity profile of the models with a CSF
layer are quite different from those without the CSF layer. However, the geometry of the sulci and the
boundary between the gray and the white matter have little effect on the detected light distribution.
© 1997 Optical Society of America
Key words: Near-infrared spectroscopy, optical path length, spatial sensitivity profile, oxygenation
monitoring.
1. Introduction
Since its first proposal
1
the technique of near-
infrared spectroscopy ~NIRS! has been increasingly
applied for the noninvasive measurement of tissue
oxygenation in the brain,
2–6
and several different in-
struments are now available for clinical
monitoring.
7–10
The development of the quantita-
tive measurement of absorption change by a modified
Beer–Lambert law made a significant advance in
NIRS studies.
11
The quantification of NIRS data re-
quires a knowledge of the optical path length in the
tissue, which is considerably farther than the physi-
cal distance between source and detector. Direct
time-of-flight measurement with a picosecond pulsed
laser and streak camera initially enabled the mean
flight time ~^t&!, and hence the mean optical path
length could be derived experimentally for a rat
head,
11
an adult head,
12
and a neonatal head
12,13
;
larger studies have recently been completed in which
phase-resolved techniques were used.
14,15
In the
NIRS calculations, the head is assumed to be a ho-
mogeneous medium, although in reality the source
and detection fibers are attached onto the surface of
the head, requiring the light to pass through the
surface tissue layers such as scalp, skull, and cere-
brospinal fluid ~CSF! both before and after passing
through the brain tissue. The clinically important
factors in NIRS monitoring of cerebral oxygenation
are the contribution of the absorption change in the
brain to the detected signal and the volume of tissue
interrogated, and these are obviously affected by the
inhomogeneity of the head and the measurement ge-
ometry. Because these factors cannot be obtained
experimentally, it is vital to be able to predict accu-
rately the light propagation in an inhomogeneous
structure, such as the head, by mathematical meth-
ods.
Several different mathematical techniques have
When this work was performed, E. Okada, M. Firbank, M.
Schweiger, M. Cope, and D. T. Delpy were with the Department of
Medical Physics and Bioengineering, University College London,
First Floor, Shropshire House, 11-20 Capper Street, London WC1E
6JA, UK. S. R. Arridge was with the Department of Computer
Science, University College London, Gower Street, London WC1
6BT, UK. E. Okada is now with the Department of Electronics
and Electrical Engineering, Keio University, 3-14-1 Hiyoshi,
Kohoku-ku, Yokohama 223, Japan.
Received 8 March 1996; revised manuscript received 12 June
1996.
0003-6935y97y00021-11$10.00y0
© 1997 Optical Society of America
1 January 1997 y Vol. 36, No. 1 y APPLIED OPTICS 21
been used to describe light propagation in scattering
tissue, and some preliminary modeling of simple lay-
ered structures have shown that the light penetra-
tion into deeper regions ~e.g., the brain! is strongly
affected by the optical properties of the surface
layer.
16–24
The presence of a relatively clear layer
~e.g., CSF! that has both low scattering and absorp-
tion coefficients has been shown especially to alter the
light propagation in the head.
21,24
However, in al-
most all these studies, the boundary of each layer has
had a simple geometry such as a flat or curved sur-
face that is significantly different from real head
structures. More sophisticated models are needed
for a rigorous analysis of light propagation in the
adult head. For example, the brain surface is actu-
ally deeply folded with many CSF-filled sulci, and it
is likely that light propagation in the brain is affected
by the sulcus structure.
In this study, in a variety of models of the adult
head, the effect of both the presence of the surface
tissues ~including a clear CSF layer! around the brain
and the brain anatomy itself have been investigated
by the use of solid slab phantoms that consist of lay-
ers with different optical properties. The simplest
adult head model is a three-layered slab without a
clear layer, whereas the most sophisticated model
has four layers, including a clear layer together with
slots that imitate the sulci. Time point-spread func-
tions for several different detection positions on the
outer surface are measured with a picosecond laser
and a streak camera, and the effect of the layered
structures is evaluated in terms of the mean optical
path length. The experimental data are compared
with the results of both Monte Carlo ~MC! and finite-
element calculations, which are also used to predict
the mean optical path length in each layer. Because
the path of individual photons can be traced in the
MC calculation, the spatial sensitivity profile in each
model has also been predicted, and the effects of the
surface layers on the volume of tissue interrogated in
the adult brain by NIRS instruments are discussed.
2. Optical Path in Near-Infrared Spectroscopy
The NIRS technique relies on the application of a
modified Beer–Lambert law
11
to convert measured
variations in attenuation ~DOD, where OD is the op-
tical density! into quantitative changes in the absorp-
tion coefficient ~Dm
a
! in the tissue. In the modified
Beer–Lambert law, the mean optical path length ^L&
replaces the physical distance between the source
and the detector:
DOD < Dm
a
^L& 5 Dm
a
c^t&.(1)
The mean optical path length is significantly greater
than the distance between the source and the detec-
tor because of the large amount of scattering in the
tissue. Thus a priori knowledge of the mean optical
path length is needed to quantify the change in ab-
sorption by the use of the modified Beer–Lambert
law. The mean optical path length can be derived
from the mean time of flight ^t& and the speed of light
c in the tissue. The mean time of flight can be ob-
tained from the temporal point-spread function
~TPSF!, which is the temporal intensity distribution
of a picosecond pulsed light that is broadened because
of the different scattering paths in the medium.
Although biological tissue has an inhomogeneous
structure, the tissue is assumed to be homogeneous
in the modified Beer–Lambert law. If it can be as-
sumed that the inhomogeneous tissue consists of sev-
eral homogeneous media, a partial mean optical path
length can be defined.
19
The partial mean optical
path length ^L
i
& is the mean optical path length that
the detected light travels within a particular medium
i, and the variation in attenuation of the detected
light across the tissue can be approximated by the
sum of the product of the partial mean optical path
lengths and the corresponding absorption coefficient
changes in each layer ~Dm
ai
!:
DOD <
(
Dm
ai
^L
i
&.(2)
The mean optical path length ^L&, which is the sum
of the partial mean optical path length ^L
i
&, can be
calculated from the experimental TPSF measured
with a picosecond pulsed laser system. However,
the TPSF contains no direct information about the
time that the light has spent in each medium, and
hence the partial mean optical path length cannot be
obtained experimentally.
If the change in the absorption coefficient in each
medium is the same, then the partial mean optical
path length indicates the contribution that each me-
dium makes to the change in the output signal. It
should, however, be noted that, in general, the
changes in absorption in each medium are not the
same because of differences in hemoglobin content.
For instance, the normal CSF layer contains no he-
moglobin, and hence it does not contribute to the
change in the output signal. Although the partial
mean optical path length indicates the signal contri-
bution of each medium, it does not show the spatial
distribution of the volume contributing to the output
signal. The volume of tissue interrogated with
NIRS instruments can, however, be calculated as the
spatial sensitivity profile,
22
which is deduced from
the accumulated optical path histories of the photons
reaching the detector.
25,26
3. Methods
A. Adult Head Models
The adult head models used in this study are inho-
mogeneous slabs that consist of three or four different
homogeneous media. The geometries of the models
and the optical properties for each layer of the model
are shown in Fig. 1 and Table 1, respectively. The
simplest three-layered model @Fig. 1~a!# consists of a
12-mm-thick surface layer that imitates the scalp
and skull, a 4-mm-thick gray-matter layer, and a
white-matter layer. The optical properties for these
layers have been chosen from the reported data on
the optical properties of tissue.
27–29
The first four-
22 APPLIED OPTICS y Vol. 36, No. 1 y 1 January 1997
layered model @Fig. 1~b!# has a 2-mm-thick clear layer
that imitates the CSF between the 10-mm-thick sur-
face layer and the gray-matter layer. The simplest
model of the brain structure @Fig. 1~c!# has an uneven
boundary between the gray matter and the white
matter. The gray-matter layer has thick areas, 14
mm in thickness and 9 mm in width, placed every 15
mm. In the most sophisticated brain model @Fig.
1~d!#, slots 10 mm deep and 1 mm wide that imitate
the sulci filled with the clear CSF are added to the
simple brain model. The thickness of each layer and
the geometry of the sulci and gray matter were cho-
sen from a magnetic resonance image of an adult
head.
B. Experimental Setup
The actual design of the adult head phantoms is
shown in Fig. 2. The phantom was made of epoxy
resin containing TiO
2
to alter its scattering coefficient
and IR absorbing dyes to alter its absorption coeffi-
cient.
30
The phantom for each model consisted of
two parts. The surface layer of 10 mm and 12 mm in
thickness ~which imitated the scalp and skull! formed
the front and the rear walls, respectively, of a box
filled with glycerol, which imitated the CSF. The
second part, a block, 13 cm wide 3 6 cm high 3 8cm
thick, which consisted of the gray-matter and the
white-matter layers, could be positioned inside the
box at a suitable distance away from the inner face of
the front or rear wall. Three different inner blocks
were made with gray- and white-matter geometries,
as shown in Fig. 1. The three-layered model @Fig.
1~a!# was realized when the inner slab was positioned
directly against the inner face of the 12-mm surface
layer. The inner slab was located 2 mm away from
the inner face of the 10-mm surface layer for all the
other models. A picosecond pulsed laser and streak
camera were used to measure the TPSF of the phan-
toms.
11
The laser system consisted of an Ar-ion la-
ser pumping a Ti:sapphire laser and streak camera.
Laser pulses of approximately 2-ps half-maximum
width at the 800-nm wavelength were emitted at 82
MHz. Most of the laser light was delivered to the
surface of the phantom while a part of the laser beam
was sampled and directly relayed to the streak cam-
era as a time-reference pulse. In the case of the
simplest brain model @Fig. 1~c!#, the irradiated spot
was just over the center of the thick gray matter, and
in the sophisticated brain model @Fig. 1~d!#, it was in
the same position, which now coincided with the cen-
ter of a slot. The light emerging from the phantom
was collected in a fiber bundle and was conveyed to
the streak camera. The spacing between the irradi-
ated spot and the fiber bundle was altered horizon-
tally for all the models and also vertically in the case
of the models of Figs. 1~c! and 1~d!~i.e., along the
thick area of the gray matter or the slot!. The
TPSF’s at each source–detector spacing were mea-
sured and stored on a computer. The mean time of
flight was calculated from the TPSF, following soft-
Fig. 1. Schematic designs of the adult head models.
Fig. 2. Construction details for the adult head phantoms.
Table 1. Optical Properties of the Adult Head Models
Tissue Type
Transport Scattering
Coefficient m
s
9
~mm
21
!
Absorption
Coefficient
m
a
~mm
21
!
Scalp and skull 2.0 0.04
CSF 0.01 0.001
Gray matter 2.5 0.025
White matter 6.0 0.005
1 January 1997 y Vol. 36, No. 1 y APPLIED OPTICS 23
ware corrections for nonlinearity, shading sensitivity,
etc., of the streak camera.
C. Monte Carlo Simulation
The MC algorithm used in this study has already
been described
19
and is based on the variance-
reduction technique.
31–33
Isotropic scattering was
assumed, and if a photon crossed the boundary be-
tween different media, the distance to the next scat-
tering event was corrected by the use of the transport
scattering coefficient in the subsequent medium m
sj
9:
l
j
5 ~l
i
2 Dl!m
si
9ym
sj
9,(3)
where l
j
is the path length to the next scattering from
the boundary of the media and Dl is the path length
to the boundary of the media from the previous scat-
tering point. In order to avoid dividing by zero in
this correction process, a transport scattering coeffi-
cient of 0.01 was used for the clear CSF layer. Re-
flection and refraction of light caused by refractive-
index mismatching between the air and the tissue
were taken into account.
When the photon was scattered out of the head
model, the survival weight of the photon was calcu-
lated from the absorption coefficients m
ai
and the ac-
cumulated partial optical path length in each
medium L
i
. The survival weight and partial path
length of the photon were recorded for each detection
position up to a distance of 65 mm from the source.
For cases in which the photon reached detection po-
sitions of 15, 30, or 40 mm, the history of the photon
path weighted by the survival weight was accumu-
lated to obtain the spatial sensitivity profiles.
22
The
photon paths were projected onto an xz plane to
record the two-dimensional spatial sensitivity pro-
files. After 10,000,000 input photons were traced,
the mean optical path length, partial mean optical
path length, the intensity of the detected light nor-
malized by source intensity, and the spatial sensitiv-
ity profiles were calculated.
D. Finite-Element Method
The time-independent diffusion equation,
34
which is a
well-known approximation of the radiative transfer
equation,
35
has been used to describe light propaga-
tion in tissue:
z k~r!¹F~r! 1 m
a
~r!F~r! 5 q
0
~r!,(4)
where k~r! is the diffusion coefficient, k~r!5$3@m
a
~r!
1m
s
9~r!#%
21
, F~r! is the photon density, and q
0
~r! is
the isotropic source distribution. In this study, a
finite-element method
36,37
~FEM! was used to solve
the diffusion equation, and the outgoing fluence ~exi-
tance!G~r!was calculated by
G~r! 5 2k~r!eˆ z ¹F~r!,(5)
where eˆ is the vector normal to the detection area.
In order to obtain the partial mean optical path
length from the exitance, it is assumed that the op-
tical path length in each layer does not vary with a
small absorption change. The difference in exitance
DG caused by a 1% absorption change in a particular
layer i was predicted, and the partial mean optical
pathlength ^L
i
& in layer i was calculated by
^L
i
& 5 DGyDm
ai
.(6)
Although the FEM can be applied to three-
dimensional models, in this study a two-dimensional
rectangular model was used to keep the memory size
required for matrix manipulation within a reason-
able limit. For the three-layered model, the rectan-
gular domain was divided into approximately 21,000
triangular subspaces and Robin boundary conditions
were used.
37
An isotropic point source located at a
distance 1ym
s
9 below the surface layer at the irradi-
ated position was used to approximate a collimated
incident laser beam.
37
Because the diffusion equation no longer holds in a
medium that has a low scattering coefficient, the
FEM could not be directly applied to the models of
Figs. 1~b!–1~d! with a clear CSF layer, so a hybrid
radiosity–diffusion theory model
24
was used instead.
The concept of the hybrid radiosity–diffusion model is
to predict light propagation in the scattering regions
by the diffusion theory and in the clear CSF layer by
a radiosity method and to combine the two results in
an iterative scheme until a minimum change in exi-
tance is achieved. It is assumed that the light in the
scattering tissue, such as scalp, skull, and brain,
obeys the diffusion equation ~4!, and the light propa-
gating without diffusion in the CSF complies with the
radiosity equation.
24
The FEM was applied to the
two rectangular domains, one being the surface layer
and the other the gray- and the white-matter layers.
The surface and the inner domains were divided into
approximately 7000 and 13,000 triangles, respec-
tively. The photon density at the inner boundary of
the surface domain arising from the incident beam on
its outer surface was first calculated by the FEM, and
from this resulting photon density the outgoing radi-
ance I~ p, s! at any point p traveling in direction s on
the surface was obtained. With this outgoing radi-
ance, the radiosity theory was used to calculate the
resulting irradiance G
R
~q! on an area dA at position q
on the outer surface of the inner domain across the
clear CSF layer:
G
R
~q! 5
**
A
I~p, s!cos~u
1
!cos~u
2
!
uru
2
exp~2urum
a
!dA,(7)
where u
1
, u
2
are the angles between s and normal
components of the surface of the two layers, m
a
is the
absorption coefficient of the CSF layer, s points from
p to q. The photon density in the inner domain
caused by the irradiance on its surface was then cal-
culated by the FEM, and the resulting outgoing ra-
diance from the inner domain was similarly obtained.
The resulting radiance was then again used as an
input to the radiosity equation, which calculated the
fluence back onto the inner face of the surface do-
main, and the photon density in the surface layer was
recalculated. This process was iterated until the
24 APPLIED OPTICS y Vol. 36, No. 1 y 1 January 1997
change in the total exitance became negligible ~,1%
change!.
The simple and the sophisticated brain models
@Figs. 1~c! and 1~d!, respectively# have different cross
sections in the xz and the yz planes, but the cross
section in the xz plane was used for the two-
dimensional approximation for both these models.
The partial mean optical path lengths in the models
with the clear CSF layers were calculated in the same
way as for the three-layered model.
4. Results
Experimental results for the mean time of flight are
compared with the predictions of the MC method
~MCM! and the FEM in Fig. 3. The corresponding
mean optical path length calculated from the mean
time and the speed of light in the epoxy resin is also
shown in each figure. On a SunSparc 20 worksta-
tion, the calculation for each model took ;200-h CPU
time for the MCM, 3-min CPU time for the FEM
based on the diffusion theory, and 15-min CPU time
for the FEM based on the hybrid radiosity–diffusion
theory. The experimental results and predictions
~MCM and FEM! for the mean time of flight as a
function of spacing are in good agreement for all mod-
els. However, statistical noise in the MC results is
notable at detection positions of greater than 30 mm.
The mean time for all the models at detection posi-
tions up to 20 mm are almost the same. For the
models of Figs. 1~b!–1~d!, which have the clear CSF
layer, once the detection position is greater than 20
mm the mean time increases only slowly with spacing
between source and detector whereas for the three-
layered model @Fig. 1~a!# without the clear layer, it
continues to increase rapidly. The differences in
mean time between all the models with a clear CSF
layer are not significant over the whole range of de-
tection positions. In both the simple brain model
@Fig. 1~c!# and the sophisticated brain model @Fig.
1~d!#, the direction of the detection position ~horizon-
tal or vertical! produced no statistically significant
differences in mean time, so data in regard to the
vertical detection positions are not shown.
The partial mean optical path length in each layer
in the models is shown in Fig. 4. The partial optical
path length cannot be obtained from the experimen-
tal TPSF, so only the results predicted by the MCM
and the FEM are compared. The predictions of both
methods show the same tendency. In the three-
layered model the partial mean optical path length of
both the gray- and the white-matter layers is small at
detection positions up to 30 mm. This means that
Fig. 3. Mean time of flight and corresponding mean optical path length as functions of the detection position predicted by the MCM and
the FEM compared with experimental results ~61SD!.
1 January 1997 y Vol. 36, No. 1 y APPLIED OPTICS 25
the light is largely confined to the surface layer.
Once the detection position is greater than 30 mm,
the partial mean optical path lengths of both the
gray- and the white-matter layers ~and beyond 50
mm especially the white-matter layer! steeply in-
crease. From the FEM results, the partial mean
optical path length of the white-matter layer exceeds
that of the surface layer at detection positions of
greater than 60 mm. In the models of Figs. 1~b!–1~d!
with the clear CSF layer, the relationships between
detection position and partial mean optical path
lengths are similar, but are completely different from
those in the three-layered model of Fig. 1~a!. The
partial mean optical path lengths of the deeper layers
are small at detection positions up to 15 mm. Once
the detection position exceeds 15 mm, the partial
mean optical path lengths of both the CSF and the
gray-matter layer start to increase. The partial
mean optical path lengths of both the surface and the
gray-matter layers remain almost constant at detec-
tion positions of greater than 30 mm, whereas that of
CSF layer continues to gradually increases. The
partial mean optical path length of the white-matter
layer is still small, even when the detection position
exceeds 60 mm.
The intensity of detected light predicted by both
the MCM and the FEM is shown in Fig. 5. The
results are normalized by the source intensity, and in
the case of the simple and the sophisticated brain
models @Figs. 1~c! and 1~d!, respectively# only the re-
sults of horizontal detection are shown. The MCM
and FEM predictions agree well for all the models.
Up to 20 mm, the intensity for all the models is the
same as a function of the detection position. Beyond
20 mm, the rate of decline in the intensity with the
detection position for all the models with a clear CSF
layer diminishes whereas that for the three-layered
model continues at approximately the same rate.
The MC-calculated spatial sensitivity profiles for
the three- and the four-layered models @Figs. 1~a! and
1~b!, respectively# at a detection position of 15 mm are
shown in Figs. 6~a! and 6~b!, respectively. The con-
tours are drawn for every 12.5% fall from the maxi-
mum sensitivity point, and the extreme contour
indicates a relative sensitivity of 0.3%. In both Figs.
6~a! and 6~b! the spatial sensitivity profiles are con-
Fig. 4. Partial mean optical path length as a function of detection position predicted by the MCM ~symbols! and the FEM ~curves!.
26 APPLIED OPTICS y Vol. 36, No. 1 y 1 January 1997
fined to the surface layer, so the results for Fig. 6~b!
will also apply to the simple and the sophisticated
brain models of Figs. 1~c! and 1~d!, respectively,
whose differences in the geometry occur only under
the CSF layer. Figures 7~a!–7~d! show the spatial
sensitivity profiles for all the models at a detection
position of 30 mm. In the case of the simple and the
sophisticated brain models, the results of horizontal
detection are also shown. At this detection position
the number of detected photons is not sufficient to
provide good statistics; however, the general ten-
dency for the localization of the sensitive area can be
recognized. In the three-layered model of Fig. 7~a!
the spatial sensitivity profile is still largely confined
to the surface layer, and little light penetrates into
the gray matter. The spatial sensitivity profile of
the four-layered model of Fig. 7~b! spreads farther
toward the clear CSF and the gray-matter layers;
Fig. 5. Normalized intensity of detected light for each model as a
function of the detection position predicted by the MCM and the
FEM.
Fig. 6. Spatial sensitivity pro-
files with a detection fiber 15 mm
distant from the light source
along the horizontal ~x! axis.
Fig. 7. Spatial sensitivity pro-
files with a detection fiber 30 mm
distant from the light source
along the horizontal ~x! axis.
1 January 1997 y Vol. 36, No. 1 y APPLIED OPTICS 27
however, very little light reaches the white-matter
layer. Although light reaches the gray-matter layer,
there are no significant differences between the spa-
tial sensitivity profiles of all three models with a clear
CSF layer @Figs. 7~b!–7~d!#. The spatial sensitivity
profiles at a detection position of 40 mm are shown in
Figs. 8~a! and 8~d!. Contours have not been drawn
on these profiles because of poor statistics. In the
three-layered model of Fig. 8~a! the spatial sensitivity
profile is still confined mainly to the surface layer.
In the other models @Figs. 8~b!–8~d!# the spatial sen-
sitivity profile has shifted toward the deeper layers,
and an apparent light path can be seen around the
clear CSF layer. However, the detected light still
does not tend to penetrate into the white-matter
layer. A slight difference in the spatial sensitivity
profiles can be observed between the four-layered
model of Fig. 8~b!, the simple brain model of Fig. 8~c!,
and the most sophisticated brain model of Fig. 8~d!.
Finally, Fig. 9 shows the vertical spatial sensitivity
profiles of all the models at a detection position of 30
mm. The spatial sensitivity profiles for all the mod-
els with a clear CSF layer spread around the CSF
layer. The profiles of the four-layered model of Fig.
9~b! and the simple brain model of Fig. 9~c! are almost
identical. In the most sophisticated brain model of
Fig. 9~d!, the source and the detector were positioned
above a sulcus, and it can be seen that light pene-
trates more deeply in the area along the sulcus.
However, the spatial sensitivity profile around the
sulcus is still confined to the gray-matter layer with
little penetration into the white matter.
5. Discussion
In the adult head, experiment has shown that the
mean optical path length divided by the spacing be-
tween source and detector, the differential path-
length factor, is approximately constant for detection
positions greater than 25 mm
12
at a value of ;6, but
increases at closer detection positions. Because the
mean optical path lengths for all the phantoms with
a clear CSF layer also show this tendency, the results
from these models are thought to reasonably mimic
the actual light propagation in the adult head.
In the MCM, the full three-dimensional geometry
is faithfully replicated, and therefore the errors in
prediction are caused mainly by inadequate photon
statistics. The MC results show that although the
clear CSF layer increases the intensity of the de-
tected light and hence improves the statistics, signif-
icant error is still notable once the detection position
is beyond 30 mm. On the other hand, both the nor-
mal and the hybrid FEM predictions show reasonable
agreement with the experimental results in spite of
being only two-dimensional approximations. This
indicates that these techniques can probably be used
with some confidence to calculate mean optical path
lengths in complex heterogeneous media. This is
important because, in the practical use of NIRS on
the human adult, the fiber spacing is often from 30 to
60 mm, and the statistical error of MC predictions for
reasonable computation times is large, beyond a de-
tection position of 30 mm.
The mean optical path lengths for the models with
Fig. 8. Spatial sensitivity pro-
files with a detection fiber 40 mm
distant from the light source
along the horizontal ~x! axis.
28 APPLIED OPTICS y Vol. 36, No. 1 y 1 January 1997
a clear CSF layer increase slowly beyond a detection
position of 25 mm whereas those of the three-layered
model continue to increase steeply. It is obvious
that the clear CSF layer considerably affects the
mean optical path length at these large spacing.
The features of light propagation in the adult head
can thus be placed in these categories according to
the detection position: ~1! at small detection posi-
tions ~#15 mm! the mean optical path length is
equivalent to the partial mean optical path length of
the surface layer, i.e., the spatial sensitivity profile is
confined to the surface layer; ~2! at intermediate de-
tection positions ~$15 mm, #25 mm! the partial
mean optical path lengths of both the clear CSF and
the gray-matter layer increase with the detection po-
sition, and the spatial sensitivity profile spreads lat-
erally over the inner face of the surface layer and the
gray-matter layer; and ~3! at large detection positions
~$25 mm! the partial mean optical path lengths of
the surface and the gray-matter layer remain approx-
imately constant while that of the clear CSF layer
increases with the detection position. The spatial
sensitivity profile is distributed mainly around the
surfaces that face the clear CSF layer except for the
sites directly underneath the source and the detector.
It is apparent that once light reaches the clear CSF,
this layer starts to act as a conduit for the light that
reaches the distant detector. Because little absorp-
tion occurs in the clear CSF layer, the intensity of
detected light at large detection positions in the mod-
els with a clear CSF layer is much higher than that in
the three-layered model without the clear CSF layer.
The light that passes through the CSF layer only
grazes the surface and the gray-matter layers, which
therefore make little contribution to the optical path
length. Thus the partial mean optical path lengths
in the surface and the gray-matter layers increase
only slowly with detection position. It is also nota-
ble that the presence of the clear CSF layer signifi-
cantly reduces the light penetration into the deeper
white-matter layer. As shown in Fig. 4, the partial
mean optical path length of the white-matter layer in
the models with a clear CSF layer are almost negli-
gible whereas that in the three-layered model with-
out the clear CSF layer steeply increases beyond a
detection spacing of 40 mm. It should be empha-
sized here that the presence of the clear layer does
not prevent light penetration into white matter. In-
deed, the intensity of the detected light that has pen-
etrated into the white matter is probably similar to
that in the three-layered model; however, the in-
creased intensity of the light guided to the detector
through the CSF layer becomes the dominant com-
ponent of the signal. Because in the clear-layer
models, light penetration into the gray matter is pre-
dominantly confined to a shallow depth, the geometry
of both the sulci and the boundary between the gray-
and the white-matter layers only slightly affect the
partial mean optical path lengths and spatial sensi-
tivity profiles. In the case in which both the source
and the detector are placed along the axis of a slot,
the light does penetrate deeper, but even then, it is
still confined to the area around the sulcus and the
light does not penetrate the white matter.
From the FEM predictions of the partial optical
path lengths in the sophisticated brain model shown
in Fig. 4~d!, light detected at a spacing of 50 mm
spends approximately 65% of its path length in the
scalp and skull, 35% in the CSF, and 5% in the gray
matter, with very little white-matter component.
However, these ratios do not necessarily represent
the contribution of each layer to the change in the
output signal. The change in the output signal de-
pends not only on the partial mean optical path
length but also on the change in the absorption coef-
ficient. Any layer in which no absorption change
occurs does not contribute to the output signal, no
matter how long the light path length in it. Simi-
larly, the deeper area in which the partial mean op-
tical path length is much shorter than in the shallow
area can contribute significantly to the output signal
if the absorption change there is much greater than
that in the shallow area. This may explain why ab-
sorption changes caused by intracranial hemorrhage
can be detected with NIRS instruments even if they
occur in the white matter or in deeper areas of the
gray matter.
38
This sort of drastic absorption
change will also considerably affect the light propa-
gation in the head and hence alter both the partial
mean optical path lengths and the spatial sensitivity
profile. For example, it is easy to imagine that if the
CSF is replaced by blood in an epidural hemorrhage,
Fig. 9. Spatial sensitivity profiles with a detection fiber 30 mm
distant from the light source along the vertical ~ y! axis.
1 January 1997 y Vol. 36, No. 1 y APPLIED OPTICS 29
the CSF layer no longer works as a light guide. Ac-
cordingly the prediction of partial mean optical path
length and spatial sensitivity profile for the models in
this study cannot be used to analyze the signal con-
tribution and interrogated area in cases in which
drastic absorption changes occur in the head.
However, in most clinical NIRS studies, smaller
absorption changes, for example, those that are due
to mild hypoxia or changes in oxygenation state with
brain activity and so on, are normally monitored.
This sort of small absorption change should mini-
mally affect the light propagation in the head, and
hence the detected light distribution can be described
by the spatial sensitivity profiles shown in Figs. 6–9.
Because the change in the absorption coefficient de-
pends on the blood content and there is virtually no
absorption change in the CSF layer under normal
conditions, the contribution of the CSF layer to the
output signal is negligible. In this study the scalp
~typically 5 mm thick! and the skull ~typically 5 mm
thick! were combined into one surface layer, and the
partial optical path length of this combined layer is
much greater than that of all other layers. The
blood volume and hence the absorption coefficient of
the scalp is much higher than those of the skull;
however, in NIRS studies, the blood directly under
the NIRS optodes may often be squeezed out because
of optode pressure, thus reducing its contribution to
the total absorption change signal. The light in the
surface layer then passes through the skull, which,
because of its low blood volume, also contributes little
to any absorption change signal. This supposition is
further borne out by the spatial sensitivity profile,
which shows that this element of the signal arises
mainly from the inner skull table. In the brain the
blood volume in gray matter is approximately twice
as much as that in white matter,
39
and hence the
absorption change caused by oxygenation variation
in gray matter is generally greater than that in white
matter. Consequently, the actual contribution of
the absorption change in the gray matter to the
change in the detected NIRS signal is probably sig-
nificantly greater under normal conditions and it
probably reaches at least 20–30%. The success of
experimental studies of cerebral-evoked response in
adult humans provides further evidence for this
conclusion.
6,4042
6. Conclusions
In this study the effect on NIR signals of the layered
surface tissues surrounding the brain in the adult
head have been investigated by both time-of-flight
measurements and mathematical predictions. The
clear CSF layer significantly affects light propagation
in the adult head once the spacing between source
and detector is greater than 15 mm. At large
source–detector spacing the detected light passes
mainly through the CSF layer and this forces the
sensitive region to be confined to a shallow section of
the gray matter. The partial mean optical path
lengths of both the surface and the gray-matter lay-
ers change only slowly once the detection position
exceeds 30 mm. This indicates that the contribution
of changes in absorption in the gray matter to the
NIRS signal is almost constant at these detection
positions and its contribution probably reaches 20%–
30%. Under these circumstances, it is difficult for
NIRS to detect small oxygenation changes in the
deeper areas of the gray and the white matter. The
geometries of the sulci and the boundary between the
gray- and the white-matter layers scarcely affect the
optical path in the adult head.
This work was supported by the Japan Society for
the Promotion of Science, Postdoctoral Fellowship for
Research Abroad to E. Okada from April 1995 to
March 1996, funding from the Engineering and Phys-
ical Research Council ~UK!~GRyK07386, GRy
G05100!, the Wellcome Trust, and Hamamatsu
Photonics KK. Some parts of these data were pre-
sented at a Society of Photo-Optical Instrumentation
Engineers’ Conference, “Photon Propagation in Tis-
sues: Quantitation and Clinical Studies using Con-
tinuous Wave, Time, and Frequency Domain
Technology” in Barcelona, Spain, in 1995.
References
1. F. F. Jo¨bsis, “Non invasive, infrared monitoring of cerebral and
myocardial oxygen sufficiency and circulatory parameters,”
Science 198, 1264–1267 ~1977!.
2. C. J. Aldrich, J. S. Wyatt, J. A. Spencer, E. O. R. Reynolds, and
D. T. Delpy, “The effect of maternal oxygen administration of
human fetal cerebral oxygenation measured during labour by
near infrared spectroscopy,” Br. J. Obstet. Gynaecol. 101, 509
513 ~1994!.
3. J. S. Wyatt, D. T. Delpy, M. Cope, S. Wray, and E. O. R.
Reynolds, “Quantification of cerebral oxygenation and haemo-
dynamics in sick newborn infants by near infrared spectrosco-
py,” Lancet 8515, 1063–1066 ~1986!.
4. M. Ferrari, E. Zanette, I. Giannini, G. Sideri, C. Fieschi, and A.
Carpi, “Effect of carotid artery compression test on regional
cerebral blood volume, haemoglobin oxygen saturation and
cytochrome-c-oxidase redox level in cerebrovascular patients,”
Adv. Exp. Med. Biol. 200, 213–222 ~1986!.
5. N. B. Hampson, E. M. Camporesi, B. W. Stolp, R. E. Moon, J. E.
Shook, J. A. Griebel, and C. A. Piantadosi, “Cerebral oxygen
availability by NIR spectroscopy during transient hypoxia in
humans,” J. Appl. Physiol. 69, 907–913 ~1990!.
6. Y. Hoshi and M. Tamura, “Detection of dynamic changes in
cerebral oxygenation coupled to neuronal function during men-
tal work in man,” Neurosci. Lett. 150, 5–8 ~1993!.
7. Y. A. B. D. Wickramasinghe, J. A. Crowe, and P. Rolfe, “Laser
source and detector with single processor for a near infra-red med-
ical application,” in IERE Progress Reports on Electronics in Med-
icine and Biology, K. Copeland, ed. ~Institution of Electronic and
Radio Engineers, London, 1986!, pp. 209–215.
8. M. Cope and D. T. Delpy, “System for long term measurement
of cerebral blood and tissue oxygenation on newborn infants by
near infrared transillumination,” Med. Biol. Eng. Comput. 26,
289–294 ~1988!.
9. B. Chance, M. Maris, J. Sorge, and M. Z. Zhang, “A phase
modulation system for dual wavelength difference spectros-
copy of haemoglobin deoxygenation in tissue,” in Time-
Resolved Laser Spectroscopy in Biochemistry II, J. R. Lakowicz,
ed., Proc. SPIE 1204, 481–491 ~1990!.
10. M. Miwa, Y. Ueda, and B. Chance, “Development of time re-
solved spectroscopy system for quantitative noninvasive tissue
measurement,” in Optical Tomography, Photon Migration,
30 APPLIED OPTICS y Vol. 36, No. 1 y 1 January 1997
and Spectroscopy of Tissue and Model Media: Theory, Hu-
man Studies, and Instrumentation, B. Chance and R. R. Al-
fano, eds., Proc. SPIE 2389, 142–149 ~1995!.
11. D. T. Delpy, M. Cope, P. van der Zee, S. R. Arridge, S. Wray,
and J. S. Wyatt, “Estimation of optical pathlength through
tissue from direct time of flight measurement,” Phys. Med.
Biol. 33, 1433–1442 ~1988!.
12. P. van der Zee, M. Cope, S. R. Arridge, M. Essenpreis, L. A.
Potter, A. D. Edwards, J. S. Wyatt, D. C. McCormick, S. C.
Roth, E. O. R. Reynolds, and D. T. Delpy, “Experimentally
measured optical pathlengths for the adult head, calf and
forearm and the head of the newborn infant as a function of
inter optode spacing,” Adv. Exp. Med. Biol. 316, 143–153
~1992!.
13. J. S. Wyatt, M. Cope, D. T. Delpy, P. van der Zee, S. R. Arridge,
A. D. Edwards, and E. O. R. Reynolds, “Measurement of optical
pathlength for cerebral near infrared spectroscopy in newborn
infants,” Dev. Neurosci. 12, 140–144 ~1990!.
14. A. Duncan, J. H. Meek, M. Clemence, C. E. Elwell, L. Tyszc-
zuk, M. Cope, and D. T. Delpy, “Optical pathlength measure-
ments on adult head, calf and forearm and the head of the
newborn infant using phase resolved optical spectroscopy,”
Phys. Med. Biol. 40, 295–304 ~1995!.
15. A. Duncan, J. H. Meek, M. Clemence, C. E. Elwell, P. Fallon,
L. Tyszczuk, M. Cope, and D. T. Delpy, “Measurement of cra-
nial optical pathlength as a function of age using phase re-
solved optical spectroscopy,” Pediatr. Res. 39, 1–7 ~1996!.
16. R. Nossal, J. Kiefer, G. H. Weiss, R. Bonner, H. Taitelbaum,
and S. Havlin, “Photon migration in layered media,” Appl. Opt.
27, 3382–3391 ~1988!.
17. H. Taitelbaum, S. Havlin, and G. H. Weiss, “Approximate
theory of photon migration in a two-layer medium,” Appl. Opt.
28, 2245–2249 ~1989!.
18. W. Cui and L. E. Ostrander, “The relationship of surface re-
flectance measurements to optical properties of layered biolog-
ical media,” IEEE Trans. Biomed. Eng. 39, 194–201 ~1992!.
19. M. Hiraoka, M. Firbank, M. Essenpreis, M. Cope, S. R. Ar-
ridge, P. van der Zee, and D. T. Delpy, “A Monte Carlo inves-
tigation of optical pathlength in inhomogeneous tissue and its
application to near-infrared spectroscopy,” Phys. Med. Biol. 38,
1859–1876 ~1993!.
20. E. Okada, M. Firbank, M. Schweiger, S. R. Arridge, J. C.
Hebden, M. Hiraoka, and D. T. Delpy, “Experimental measure-
ments on phantoms and Monte Carlo simulation to evaluate
the effect of inhomogeneity on optical pathlength,” in Optical
Tomography, Photon Migration, and Spectroscopy of Tissue
and Model Media: Theory, Human Studies, and Instrumen-
tation, B. Chance and R. R. Alfano, eds., Proc. SPIE 2389,
174–181 ~1995!.
21. M. Firbank, M. Schweiger, and D. T. Delpy, “Investigation of
‘light piping’ through clear regions of scatter objects,” in Opti-
cal Tomography, Photon Migration, and Spectroscopy of Tissue
and Model Media: Theory, Human Studies, and Instrumen-
tation, B. Chance and R. R. Alfano, eds., Proc. SPIE 2389,
167–173 ~1995!.
22. E. Okada, M. Firbank, and D. T. Delpy, “The effect of overlying
tissue on the spatial sensitivity profile of near-infrared spec-
troscopy,” Phys. Med. Biol. 40, 2093–2108 ~1995!.
23. A. H. Hielscher, H. Liu, B. Chance, F. K. Tittel, and S. L.
Jacques, “Phase resolved reflectance spectroscopy on layered
turbid media,” in Optical Tomography, Photon Migration, and
Spectroscopy of Tissue and Model Media: Theory, Human
Studies, and Instrumentation, B. Chance and R. R. Alfano,
eds., Proc. SPIE 2389, 248–256 ~1995!.
24. M. Firbank, S. R. Arridge, M. Schweiger, and D. T. Delpy, “An
investigation of light transport through scattering bodies with
non-scattering region,” Phys. Med. Biol. 41, 767–783 ~1996!.
25. J. C. Schotland, J. C. Haselgrove, and J. S. Leigh, “Photon
hitting density,” Appl. Opt. 32, 448453 ~1993!.
26. S. R. Arridge, “Photon-measurement density functions. Part
I: analytical forms,” Appl. Opt. 34, 7395–7409 ~1995!.
27. W. F. Cheong, S. A. Prahl, and A. J. Welch, “A review of the
optical properties of biological tissues,” IEEE J. Quantum
Electron. 26, 2166–2185 ~1990!.
28. P. van der Zee, M. Essenpreis, and D. T. Delpy, “Optical prop-
erties of brain tissue,” in Photon Migration and Imaging in
Random Media and Tissues, R. R. Alfano and B. Chance, eds.,
Proc. SPIE 1888, 454465 ~1993!.
29. M. Firbank, M. Hiraoka, M. Essenpreis, and D. T. Delpy,
“Measurement of the optical properties of the skull in the
wavelength range 650–950 nm,” Phys. Med. Biol. 38, 503–510
~1993!.
30. M. Firbank and D. T. Delpy, “A design for a stable and repro-
ducible phantom for use in near infra-red imaging and spec-
troscopy,” Phys. Med. Biol. 38, 847–853 ~1993!.
31. B. C. Wilson and G. Adam, “A Monte Carlo model for the
absorption and flux distributions of light in tissue,” Med. Phys.
10, 824830 ~1983!.
32. P. van der Zee and D. T. Delpy, “Simulation of the point spread
function for light in tissue by a Monte Carlo technique,” Adv.
Exp. Med. Biol. 215, 179–191 ~1987!.
33. M. S. Patterson, B. C. Wilson, and D. Wyman, “The propaga-
tion of optical radiation in tissue. I. Models of radiation
transport and their application,” Lasers Med. Sci. 6, 155–168
~1990!.
34. J. J. Duderstadt and L. J. Hamilton, Nuclear Reactor Analysis
~Wiley, New York, 1978!, 140–144.
35. A. Ishimaru, Wave Propagation and Scattering in Random
Media ~Academic, New York, 1978!, Vol. 1.
36. S. R. Arridge, M. Schweiger, M. Hiraoka, and D. T. Delpy, “A
finite element approach for modeling photon transport in tis-
sue,” Med. Phys. 20, 299–309 ~1993!.
37. M. Schweiger, S. R. Arridge, M. Hiraoka, and D. T. Delpy, “The
finite element method for the propagation of light in scattering
media: boundary and source conditions,” Med. Phys. 22,
1779–1792 ~1995!.
38. S. P. Gopinath, C. S. Robertson, R. G. Grossman, and B.
Chance, “Near infrared spectroscopic localisation of intracra-
nial hematomas,” J. Neurosurg. 79, 43–47 ~1993!.
39. K. L. Leenders, D. Perani, A. A. Lammerstsma, J. D. Heather,
P. Buckingham, M. J. R. Healy, J. M. Gibbs, R. J. S. Wise, J.
Hatazawa, S. Herold, R. P. Beaney, D. J. Brooks, T. Spinks, C.
Rhodes, R. S. J. Frackowiak, and T. Jones, “Cerebral blood
flow, blood volume and oxygen utilization,” Brain 113, 27–47
~1990!.
40. A. Villringer, J. Planck, C. Hock, L. Schleinkofer, and U.
Dirnagl, “Near infrared spectroscopy ~NIRS!: a new tool to
study hemodynamic changes during activation of brain func-
tion in human adults,” Neurosci. Lett. 154, 401–404 ~1993!.
41. G. Gratton, P. M. Corballis, E. Cho, M. Fabiani, and D. C.
Hood, “Shades of gray matter: noninvasive optical images of
human brain responses during visual stimulation,” Psycho-
physiology 32, 505–509 ~1995!.
42. J. H. Meek, C. E. Elwell, M. J. Khan, J. Romaya, J. S. Wyatt,
D. T. Delpy, and S. Zeki, “Regional changes in cerebral hae-
modynamics as a result of a visual stimulus measured by near
infrared spectroscopy,” Proc. R. Soc. London Ser. B 261, 351–
356 ~1995!.
1 January 1997 y Vol. 36, No. 1 y APPLIED OPTICS 31
... A continuous-wave imaging system (NIRSport2 data acquisition system by NIRx medical technologies (Germany) was used to acquire fNIRS data. Eight emi ers and eight detectors were positioned over the motor cortex with a separation of 3 cm [30] according to the 10-20 standard system as shown in Figure 2. Twenty channels were created by the arrangement of optodes (emi er and detectors pair) on the expected motor cortex region. ...
... A continuous-wave imaging system (NIRSport2 data acquisition system by NIRx medical technologies (Germany) was used to acquire fNIRS data. Eight emitters and eight detectors were positioned over the motor cortex with a separation of 3 cm [30] according to the 10-20 standard system as shown in Figure 2. Twenty channels were created by the arrangement of optodes (emitter and detectors pair) on the expected motor cortex region. ...
... A continuous-wave imaging system (NIRSport2 data acquisition system by NIRx medical technologies (Germany) was used to acquire fNIRS data. Eight emi ers and eight detectors were positioned over the motor cortex with a separation of 3 cm [30] according to the 10-20 standard system as shown in Figure 2. Twenty channels were created by the arrangement of optodes (emi er and detectors pair) on the expected motor cortex region. The optodes are positioned on the motor cortex according to the 10-20 international system. ...
Article
Full-text available
Brain–computer interface (BCI) systems include signal acquisition, preprocessing, feature extraction, classification, and an application phase. In fNIRS-BCI systems, deep learning (DL) algorithms play a crucial role in enhancing accuracy. Unlike traditional machine learning (ML) classifiers, DL algorithms eliminate the need for manual feature extraction. DL neural networks automatically extract hidden patterns/features within a dataset to classify the data. In this study, a hand-gripping (closing and opening) two-class motor activity dataset from twenty healthy participants is acquired, and an integrated contextual gate network (ICGN) algorithm (proposed) is applied to that dataset to enhance the classification accuracy. The proposed algorithm extracts the features from the filtered data and generates the patterns based on the information from the previous cells within the network. Accordingly, classification is performed based on the similar generated patterns within the dataset. The accuracy of the proposed algorithm is compared with the long short-term memory (LSTM) and bidirectional long short-term memory (Bi-LSTM). The proposed ICGN algorithm yielded a classification accuracy of 91.23 ± 1.60%, which is significantly (p < 0.025) higher than the 84.89 ± 3.91 and 88.82 ± 1.96 achieved by LSTM and Bi-LSTM, respectively. An open access, three-class (right- and left-hand finger tapping and dominant foot tapping) dataset of 30 subjects is used to validate the proposed algorithm. The results show that ICGN can be efficiently used for the classification of two- and three-class problems in fNIRS-based BCI applications.
... The modeling of light propagation in biological tissues can be carried out using different mathematical methods [30,[41][42][43]. For scientists concerned with diagnostic and therapeutic applications, the main question is whether actual human brains with normal anatomy and accurate optical characteristics can be modeled for each tissue type [44]. While scattering and absorption are the most important interaction of light with tissue [45], the complete optical characteristics of tissue include four parameters measured at each wavelength: absorption coefficient (µ a ), scattering coefficient (µ s ), scattering anisotropy factor (g), and refractive index (n) [46]. ...
Chapter
This chapter explores the critical aspect of light penetration into the brain, which is essential for understanding the therapeutic potential of photobiomodulation (PBM). It discusses the absorption features of biological tissue components, the penetration profiles of scalp, skull, and brain tissues, and concludes with remarks on penetration depth. Section 1 introduces the concept of light penetration in the context of brain PBM, highlighting its importance for effective treatment outcomes. Section 2 delves into the absorption features of biological tissue components, explaining how various chromophores and tissue structures impact the transmission of light through the scalp, skull, and brain. Section 3 focuses on light penetration profiles of scalp and skull tissues. Data from animal studies (3.1) and human studies (3.2) are presented, with laboratory and simulation data offering a comprehensive understanding of light transmission through these tissues. Section 4 examines light penetration profiles of brain tissues, again drawing on data from both animal (4.1) and human (4.2) studies. The laboratory and simulation data provide insights into how light interacts with the brain’s complex structure and the factors that influence penetration depth. Section 5 concludes the chapter with remarks on penetration depth, emphasizing the importance of selecting appropriate wavelengths and treatment parameters to ensure optimal light penetration for effective PBM therapy. Overall, this chapter offers a thorough analysis of light penetration into the brain, equipping readers with the knowledge necessary to design and evaluate brain PBM treatments.
... The modeling of light propagation in biological tissues can be carried out using different mathematical methods [30,[41][42][43]. For scientists concerned with diagnostic and therapeutic applications, the main question is whether actual human brains with normal anatomy and accurate optical characteristics can be modeled for each tissue type [44]. While scattering and absorption are the most important interaction of light with tissue [45], the complete optical characteristics of tissue include four parameters measured at each wavelength: absorption coefficient (µ a ), scattering coefficient (µ s ), scattering anisotropy factor (g), and refractive index (n) [46]. ...
Chapter
Non-invasive delivery of photons from an external light source to the head and thence into the brain tissue is generally referred to as transcranial photobiomodulation (PBM). In this approach, light must pass through several types of tissue, such as the scalp, skull, periosteal, meningeal, subdural space, arachnoid mater, subarachnoid space, and pia mater, successively, until reaching the cortical surface. Hair can also act as a significant attenuator of light in the visible and near-infrared (NIR) wavelengths, and its barrier role should be taken into account when other parts of head (not the forehead) are irradiated.
... Realistic modelling of light propagation in the CSF is an important part in the overall model of the optical structure of the head (Okada et al 1997, Okada andDelpy 2003). Since the CSF has relatively low absorption and scattering, it offers paths along which light can travel for relatively long distances without significant loss of intensity. ...
Article
Full-text available
Objective. Diffuse optical tomography (DOT) provides a relatively convenient method for imaging haemodynamic changes related to neuronal activity on the cerebral cortex. Due to practical challenges in obtaining anatomical images of neonates, an anatomical framework is often created from an age-appropriate atlas model, which is individualized to the subject based on measurements of the head geometry. This work studies the approximation error arising from using an atlas instead of the neonate’s own anatomical model. Approach. We consider numerical simulations of frequency-domain (FD) DOT using two approaches, Monte Carlo simulations and diffusion approximation via finite element method, and observe the variation in 1) the logarithm of amplitude and phase shift measurements, and 2) the corresponding inner head sensitivities (Jacobians), due to varying segmented anatomy. Varying segmentations are sampled by registering 165 atlas models from a neonatal database to the head geometry of one individual selected as the reference model. Prior to the registration, we refine the segmentation of the cerebrospinal fluid (CSF) by separating the CSF into two physiologically plausible layers. Main results. In absolute DOT, a considerable change in the grey matter or extracerebral tissue absorption coefficient was found detectable over the anatomical variation. In difference imaging, a small local 10%-increase in brain absorption was clearly detectable in the simulated measurements over the approximation error in the Jacobians, despite the wide range of brain maturation among the registered models. Significance. Individual-level atlas models could potentially be selected within several weeks in gestational age in DOT difference imaging, if an exactly age-appropriate atlas is not available. The approximation error method could potentially be implemented to improve the accuracy of atlas-based imaging. The presented CSF segmentation algorithm could be useful also in other model-based imaging modalities. The computation of FD Jacobians is now available in the widely-used Monte Carlo eXtreme software.
... 19,32,35 Although NIRS technology was created to measure regional tissue oxygen bound to hemoglobin, CSF itself has been shown to alter the propagation of light in NIRS models. 32,[35][36][37][38] Our results support the potential for CSF (or the absence of tissue) to interfere with the rScO 2 reading, with a greater proportion of CSF included at the 2.5 cm measurement depth in infants with smaller HCs (Fig. 2), which is associated with lower rScO 2 independently of GA (Fig. 3). ...
Article
Background: Cerebral near-infrared spectroscopy is a non-invasive tool used to measure regional cerebral tissue oxygenation (rScO2) initially validated in adult and pediatric populations. Preterm neonates, vulnerable to neurologic injury, are attractive candidates for NIRS monitoring; however, normative data and the brain regions measured by the current technology have not yet been established for this population. Methods: This study's aim was to analyze continuous rScO2 readings within the first 6-72 h after birth in 60 neonates without intracerebral hemorrhage born at ≤1250 g and/or ≤30 weeks' gestational age (GA) to better understand the role of head circumference (HC) and brain regions measured. Results: Using a standardized brain MRI atlas, we determined that rScO2 in infants with smaller HCs likely measures the ventricular spaces. GA is linearly correlated, and HC is non-linearly correlated, with rScO2 readings. For HC, we infer that rScO2 is lower in infants with smaller HCs due to measuring the ventricular spaces, with values increasing in the smallest HCs as the deep cerebral structures are reached. Conclusion: Clinicians should be aware that in preterm infants with small HCs, rScO2 displayed may reflect readings from the ventricular spaces and deep cerebral tissue. Impact: Clinicians should be aware that in preterm infants with small head circumferences, cerebral near-infrared spectroscopy readings of rScO2 displayed may reflect readings from the ventricular spaces and deep cerebral tissue. This highlights the importance of rigorously re-validating technologies before extrapolating them to different populations. Standard rScO2 trajectories should only be established after determining whether the mathematical models used in NIRS equipment are appropriate in premature infants and the brain region(s) NIRS sensors captures in this population, including the influence of both gestational age and head circumference.
... Conversely, greater sensitivity to HRF is expected for SD distances greater than 10 mm (Fig. 3(B)), based on the previous findings on the partial path length of brain tissues [30][31][32]. As the SD distance increases, an increased spike noise is expected as fewer photons are detected ( Fig. 3(C)) [33,34]. Finally, a gaussian decay term (β in Eq (3)) was applied for the HRF based on the distance between the midpoint of the channel and the center of the HRF seed location (Fig. 3(D)). ...
Article
Full-text available
Optical brain monitoring, such as near-infrared spectroscopy (NIRS), has facilitated numerous brain studies, including those based on machine learning techniques. A large and diverse dataset is necessary for training machine learning algorithms to avoid overfitting a limited amount of data. However, recruiting sufficient subjects is challenging owing to time and budget constraints. Therefore, we propose an NIRS data generation algorithm that scales NIRS signal components, such as hemodynamic response function, physiological noise, and system spike noise, based on the source-detector distance to augment the training data. The experimental data were augmented with the generated NIRS data to train a convolutional neural network to classify self-paced left- and right-hand motor imagery. Augmenting the training dataset with 1000 generated data points increased the classification accuracy to 86.3 ± 4.1%, indicating a 26% increase compared with training on experimental data only. In addition, we applied Guided Gradient-weighted Class Activation Mapping (Grad-CAM) to visualize the class discriminative features of the input data. The Guided Grad-CAM heatmaps aligned well with the oxy-hemoglobin peaks during self-paced motor imagery. We concluded that the increased cerebral oxygenation, especially in the contralateral hemisphere, was the class-discriminative feature for classifying left- and right-hand motor imagery.
Article
We simulated the relationship between the diffuse reflectance and penetration depth of light in skin with respect to the source (S)–detector (D) distance using the nine-layered skin tissue model for three wavelengths in the visible range. The photon propagation trajectory is visualized for intuitive understanding. The average penetration depth and the diffuse reflectance component are graphically related with the S-D distance. The results of the penetration depth versus diffuse reflectance and comparison with multiple regression analysis indicate that changes of absorption and/or scattering in the upper dermis region are expected to be perceived selectively.
Article
A rigorous technique using a Monte Carlo model has been developed to determine the optical properties of biological tissue from goniometer and integrating sphere measurements. Using these techniques, the wavelength dependence of the phase function, g-value, absorption coefficient, scattering and reduced scattering coefficient were determined for postmortem neonate and adult human brain tissue over the wavelength range of 500 to 1000 nm. Single scattering phase functions as a function of wavelength have been measured using a goniometer system and optically thin tissue slices. Spectra for the absorption and scattering coefficients have been determined from a set of integrating sphere measurements, using a white light source and a CCD spectrometer. The integrating sphere data were analyzed using a novel Monte Carlo inversion technique, which makes use of the measured phase functions and which takes into account the effects of sample geometry and the angular dependence of specular reflection. This method overcomes some of the problems and shortfalls of the analytical techniques which employ Kubelka Munk or diffusion theory. The reduced scattering coefficients for all types of brain tissue showed a linear decrease with increasing wavelength. The wavelength dependence of the scattering coefficient and the phase function is shown to be considerable, and cannot be neglected.
Article
The relatively good transparency of biological materials in the near infrared region of the spectrum permits sufficient photon transmission through organs in situ for the monitoring of cellular events. Observations by infrared transillumination in the exposed heart and in the brain in cephalo without surgical intervention show that oxygen sufficiency for cytochrome a,a3, function, changes in tissue blood volume, and the average hemoglobin-oxyhemoglobin equilibrium can be recorded effectively and in continuous fashion for research and clinical purposes. The copper atom associated with heme a3 did not respond to anoxia and may be reduced under normoxic conditions, whereas the heme-a copper was at least partially reducible.
Article
In this study, we investigate the influence of layered tissue structures on the phase-resolved reflectance. As a particular example, we consider the affect of the skin, skull, and meninges on noninvasive blood oxygenation determination of the brain. In this case, it's important to know how accurate one can measure the absorption coefficient of the brain through the enclosing layers of different tissues. Experiments were performed on layered gelatin tissue phantoms and the results compared to diffusion theory. It is shown that when a high absorbing medium is placed on top of a low absorbing medium, the absorption coefficient of the lower layer is accessible. In the inverse case, where a low absorbing medium is placed on top of a high absorbing medium, the absorption coefficient of the underlying medium can only be determined if the differences in the absorption coefficient are small, or the top layer is very thin. Investigations on almost absorption and scattering free layers, like the cerebral fluid filled arachnoid, reveal that the determination of the absorption coefficient is barely affected by these kinds of structures.
Article
Time resolved spectroscopy of tissue makes it possible to quantify tissue hemoglobin concentrations because of the direct measurement of the optical path length for photon migration. However, the laser system is bulky and unwieldy and impractical for clinical studies. Thus, the application of the more compact and efficient phase modulation technology well known for fluorescence lifetime studies to time resolved spectroscopy of tissue offers opportunities to simplify the methodology and in addition to afford continuous readout of tissue photon propagation. This paper describes single and dual wavelength systems operating at two wavelengths in the deep red region based upon a time-sharing system. These devices have noise levels in a 2 Hz bandwidth of less than 2 ps and drifts of < 1ps/min. Applications of the noninvasive devices include measurement of hemoglobin deoxygenation in brain and hemoglobin and myoglobin deoxygenation in human skeletal muscle and animal models. Numerous applications to medical and biological problems now become available.
Article
Near-infrared spectroscopy is increasingly being used for monitoring cerebral oxygenation and haemodynamics. One current concern is the effect of the cerebrospinal fluid upon the distribution of light in the head. There are difficulties in modeling clear layers in scattering systems. The Monte Carlo model should handle clear regions accurately, but is too slow to be used for realistic geometries. The diffusion equation can be solved quickly for realistic geometries, but is only valid in scattering regions. In this paper we describe experiments carried out on a solid slab phantom to investigate the effect of clear regions. These experiments were used to examine the accuracy with which the different models described propagation through a clear layer inside a scattering object. We found that the presence of a clear layer had a significant effect upon the light distribution, which was modeled correctly by Monto Carlo techniques, but not by diffusion theory.
Article
Determination of the optical pathlength of light in tissue is important to quantitate NIRS data. However, the inhomogeneity of the illuminated tissues increases the difficulty of determining the relevant optical pathlength in the tissue. For instance, in the head, the contribution of the tissues overlying the brain to the total optical pathlength cannot be ignored in the monitoring of cerebral oxygenation with NIRS. In this study, time-of-flight measurements of an inhomogeneous phantom are carried out in the laboratory to examine the contribution of the overlying tissue to the optical pathlength. The phantom consists of two homogeneous components, the boundaries of which are two concentric cylinders. The TPSF is measured with a picosecond laser and a streak camera, and the change of TPSF with the distance between source and detection fibers is examined. The experimental TPSF and mean time of flight are compared with the results of a Monto Carlo simulation and a finite element model based on the diffusion equation. A comparison of the accuracy of prediction of the pathlength by each model is presented as a function of the spacing between source and detection fibers. The intensity photon measurement density functions in each of the cylinders were estimated from the Monte Carlo simulations. The results provide estimates for the amount of the NIRS signal arising from overlying tissues in the head.
Article
Near IR time resolved spectroscopy has been studied for quantitative determination of absorbance in highly scattering medium such as tissue. When a very narrow optical pulse is incedent into a scattering medium, the detected pulse through the medium broadens and the temporal profile is closely related to the optical property of the scattering medium. The photon migration in highly scattering medium can be described with the diffusion theory. Thus the optical property of the scattering medium can be determined by analyzing the shape of the detected tamporal profile with the diffusion equation. We have developed the time resolved spectroscopy (TRS) system based on a time correlated single photon counting technique for data acquisition and diffusion theory for data analysis. Pulsed laser diodes with two different wavelengths are used as light sources in the system. The system size is compact and it can be moved around a laboratory or hospital easily. We demonstrated its use in vivo experiments. As a result, we were able to accurately determine absorber concentrations in a highly scattering medium and the result of these in vivo experiments indicate possible use of the system for quantitative clinical studies.