RFC Review
RFC Review
RFC Review
_
m
1
z z h
1
m
1
h
1
+m
2
(z h
1
) h
1
z h
2
m
1
h
1
+m
2
(h
2
h
1
) h
2
z
+m
3
(z h
1
h
2
)
(3)
where m
3
= 0.118 M-units/m, consistent with the
mean over the United States. Since proles are up-
ward refracting, clutter power is not very sensitive to
m
3
[Gerstoft et al., 2003b].
A surface duct, schematically shown in Fig. 2c, has
also been used by Gerstoft et al. [2003b]; Rogers et al.
[2005], which includes an evaporation duct layer be-
KARIMIAN ET AL.: RFC REVIEW 5
neath the trapping layer:
M(z) = M
0
+
_
_
M
1
+c
0
_
z h
d
ln
z+z0
z0
_
z z
d
m
1
z z
d
z h
1
m
1
h
1
M
d
zh1
z
thick
h
1
z h
2
m
1
h
1
M
d
+m
3
(z h
2
) h
2
z
(4)
where c
0
= 0.13, m
1
is the slope in the mixed layer,
m
3
= 0.118 M-units/m, h
1
is the trapping layer base
height, and z
d
is the evaporation duct layer height
determined by:
z
d
=
_
h
d
1m1/c0
0 <
1
1m1/c0
< 2
2h
d
Otherwise
(5)
subject to z
d
< h
1
. h
1
= 0 simplies (4) to a
bilinear prole and h
2
= 0 implies standard at-
mosphere. z
thick
is the thickness of the inversion
layer, and h
2
= h
1
+ z
thick
. M
1
is determined by
M
1
= c
0
h
d
ln
z
d
+z0
z0
+z
d
(m
1
c
0
), and M
d
is the M-
decit of the inversion layer. Gerstoft et al. [2003b]
used an 11 parameter model for the environmental
refractivity prole: ve paremters for the vertical
structure as in (4), and six to model the range varia-
tions of the prole. They assumed that the trapping
layer height h
2
is range dependent and used principle
components of h
2
as a Markov process with respect
to range.
Most of the RFC studies including [Gerstoft et al.,
2004; Yardim et al., 2007; Vasudevan et al., 2007]
have used a four parameter surface based duct.
However, the frequency range of the validity of
a trilinear approximation to the surface duct re-
fractivity structure is arguable. As Fig. 3 demon-
strates, the trilinear approximation to complex re-
fractivity prole structures gets worse for modeling
wave propagation at higher frequencies. Propaga-
tion loss and clutter power of a measured prole
and its trilinear approximation are shown in this g-
ure. The prole is from the SPANDAR1998 dataset
(Run 07, range 50 km) measured by an instrumented
helicopter along the 150
0
where
0
is the clutter cross section per unit area and A
c
is
the area of the radar cell [Skolnik, 2008]:
A
c
= r
B
(c/2) sec ((r, m)) , (7)
with
B
the antenna pattern azimuthal beamwidth,
c the propagation speed, the pulse width, and is
the grazing angle which is a function of range and
the environmental refractivity. From this point on,
F(r, m) and (r, m) are shown as F and for sim-
plicity. Thus, the clutter power at the range r is
obtained as:
P
c
(r) =
P
t
G
2
B
c
0
sec()F
4
2(4r)
3
L
. (8)
The propagation factor F is calculated by numer-
ical solutions to the wave propagation problem (Sec-
tion 3.1). The sea surface-reectivity per unit area
0
is calculated from semi-empirical models that t
the experimental measurements to a function of sys-
tem parameters (Section 3.2).
The angle with which electromagnetic waves hit
the ocean surface varies with range. However, the
dependence of the clutter model on grazing angle
has been neglected at far distances from the radar
KARIMIAN ET AL.: RFC REVIEW 7
in [Rogers et al., 2000, 2005; Gerstoft et al., 2003b;
Kraut et al., 2004; Yardim et al., 2006, 2008; Vasude-
van et al., 2007; Wang et al., 2009; Douvenot and
Fabbro, 2010; Zhao et al., 2011]. The sec() term
also is a weak function of at low angles. Thus,
normalization of the clutter power by the power at
range r
0
yields the approximation:
P
c
(r)
P
c
(r
o
)
_
r
o
r
_
3
F
4
(r)
F
4
(r
o
)
. (9)
Rogers et al. [2000] considered the dependency of
the sea-surface reectivity with grazing angle in an
evaporation duct and concluded that
0
0
given
that the clutter cells are far enough from the radar.
They also investigated the existence of a minimum
wind speed under which radar return is not reliable
for duct height inversion. The minimum wind speed
(usually less than is 2 m/s) depends on the radar pa-
rameters and sensitivity.
To overcome the problem of uncertainty of
0
, geo-
metrical ray tracing and rank correlation was used by
Barrios [2004] for inversion of surface-based ducts.
The assumption that there is a single grazing an-
gle at each range is not always valid, especially
in strong surface-based ducts where multiple electo-
magnetic waves with dierent angles hit the surface
at each location. Karimian et al. [2011b] suggested
a clutter model that depends on all grazing angles
proportional to their relative powers:
P
c
(r) =
c
(r)F
4
(r)
_
()d
_
0,GIT
() sec()()
F
4
std
()
d,
(10)
where
c
(r) =
PtG
2
B
c
2(4r)
3
L
includes all grazing an-
gle independent terms,
0,GIT
is the sea surface re-
ectivity from the GIT model (discussed in Section
3.2), () is the relative energy of incident wavefronts
at each grazing angle obtained from a curved wave
beamformer, and F
std
() is the propagation factor
of a standard atmosphere at a range with the same
grazing angle. An analysis of the performance of dif-
ferent clutter models in RFC inversions is provided
in [Karimian et al., 2011a].
3.1. Wave propagation modeling
From the early days of wave propagation mod-
eling, a divergence arose due to the distinct dier-
ences in applications emphasizing environmental ef-
fects over terrain versus over the oceans. Due to the
advances in computer processing as well as innova-
tive mathematical techniques for numerically inten-
sive problem solving, the most popular techniques for
Radio Frequency (RF) propagation modeling have
converged such that these same methods are well
suited for both land and water propagation paths.
Since the emphasis of this paper is on the estimation
of refractive conditions over the ocean, this section
will describe only those RF propagation modeling
techniques and algorithms as they pertain to mod-
eling anomalous propagation eects on over-water
paths.
One of the rst radiowave propagation models that
took into account the eects of both evaporation
ducts and surface-based ducts was based on the tech-
niques described in [Kerr, 1951] and [Blake, 1980].
The model determines the coherent sum of the di-
rect and surface-reected elds within the optical in-
terference region, also accounting for divergence and
non-perfect reection by use of a modied Fresnel re-
ection coecient [Hitney and Richter, 1976]. Mod-
eling refractive eects is limited since within this re-
gion, the use of an eective earth radius factor is em-
ployed to account for non-standard conditions. For
diraction eects beyond the radio horizon, duct-
ing eects are based on a single mode model where
an empirical t to waveguide solutions are used to
modify Kerrs standard diraction method [Hitney,
1994].
For modeling of height-varying refractive condi-
tions, waveguide models oer a much higher delity
solution and have been in use since the early 1900s
[Budden, 1961]. Waveguide models employ normal
mode theory and are well suited when refractive con-
ditions do not change along the path. Due to the
high computational requirements for mode searches,
another caveat is that normal mode models are typ-
ically used beyond the radio horizon where far fewer
modes are needed for a solution [Pappert et al., 1992;
Hitney, 1994].
One of the more popular techniques for RF prop-
agation modeling is the parabolic equation (PE)
method, also known as the paraxial approximation
method. Originally used by Leontovich and Fock
[1946], the PE method allows for propagation con-
ditions to vary in both height and range. However,
the PE method was not in practical use until Hardin
and Tappert [1973] developed a technique called the
split-step Fourier (SSF) method, initially applied to
underwater acoustic propagation. The SSF method
took advantage of fast Fourier transforms that led to
extremely ecient numerical solutions of the PE. Ko
et al. [1983]; Skura et al. [1990] modied the under-
8 KARIMIAN ET AL.: RFC REVIEW
water acoustic SSF PE to model radiowave propaga-
tion in the troposphere. Since that time many im-
provements and mathematical techniques have been
introduced in the SSF PE algorithm for applications
to RF propagation in the troposphere. For an ex-
cellent treatise on the development of many of these
techniques, the reader is referred to Levy [2000].
Due to its eciency and accuracy the SSF PE
algorithm is now widely used in many radiowave
propagation models, including the model used here
to obtain results presented in this paper. A gen-
eral description of the SSF PE algorithm is given in
the following, with more details provided on specic
implementation of the model used here. Applying
the simple assumption of a slowly varying medium,
Maxwells equations can be reduced to the scalar two-
dimensional (Cartesian) elliptical Helmholtz equa-
tion:
2
(x, z)
x
2
+
2
(x, z)
z
2
+k
2
0
n
2
(x, z) = 0, (11)
where (x, z) is a function of the electric or magnetic
eld, depending on the polarization of the radiated
eld; and n is the refractive index of the medium
(implicitly also a function of x and z). The usual
starting point for the derivation of the PE is sub-
stituting the function (x, z) = e
jk0x
u(x, z) in (11),
then factor the result into, respectively, forward and
backward pseudo-dierential equations:
u(x, z)
x
+jk
0
_
1
1
k
2
0
2
z
2
+n
2
_
u(x, z) = 0, (12)
u(x, z)
x
+jk
0
_
1 +
1
k
2
0
2
z
2
+n
2
_
u(x, z) = 0. (13)
This substitution eectively removes the rapid phase
variation in , leaving u(x, z) a slowly varying func-
tion in range. In most PE models used for long range
radiowave tropospheric propagation, only the for-
ward propagating term (12) is solved, and the back-
ward propagating term is ignored.
Initial PE algorithms incorporated simple approx-
imations to (12), resulting in the standard PE (SPE).
The limitation with using the SPE is that it is
a narrow-angle approximation and leads to larger
errors when propagating at large angles, typically
greater than 10
z=0
+u
z=0
= 0, (14)
where the complex is given by
h,v
= jk
0
sin
_
1
h,v
1 +
h,v
_
. (15)
Here, is the grazing angle of the radiated eld at the
surface, is the Fresnel reection coecient - also
dependent on the grazing angle, and the subscripts
h and v refer to horizontal and vertical polarization
respectively. The discrete mixed Fourier transform
(DMFT) formulation provided by Dockery and Kut-
tler [1996] implements the impedance boundary con-
dition and derives the new split-step solution entirely
in the discrete domain. The DMFT method has the
added advantage that it retains numerical eciency
due to requiring only sine transforms. Further re-
nement of the DMFT was presented by Kuttler and
Janaswamy [2002] where they applied various dier-
ence formulations for (14) to arrive at an improved
DMFT algorithm, reducing much of the numerical
instabilities associated with the quantity
h,v
when
Re(
h,v
) approaches zero.
The propagation model used for the results pre-
sented in this paper implements the WAPE and
the DMFT algorithm as described in Kuttler [1999];
Dockery and Kuttler [1996]; Kuttler and Janaswamy
[2002] and is called the Advanced Propagation Model
(APM). The handling of range-varying vertical re-
fractive proles is described in [Barrios, 1992] and a
general description of the APM is provided in [Bar-
rios, 2003].
Pertinent to the RFC methodology is the accuracy
of the forward scattered eld, which is subsequently
dependent on how
h,v
is modeled. Typically, the
boundary condition is modeled such that a constant
impedance is assumed within each range step, de-
KARIMIAN ET AL.: RFC REVIEW 9
350 400
0
100
200
300
400
500
600
700
800
900
1000
M units
A
l
t
i
t
u
d
e
(
m
)
(a)
0 100 200 300
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Range (km)
(b)
A
n
g
l
e
(
d
e
g
)
Figure 4. (a) Refractivity prole of surface-based duct
used for (b) determination of grazing angles by ray trace
(solid) and nal maximum angles (dashed lines) used for
computing
h,v
.
pendent on a single grazing angle associated with
the dominant mode of propagation for the specied
refractive environment. We apply the Kircho ap-
proximation and model the sea surface boundary by
determining an eective impedance described by a
reduction, , to the smooth surface Fresnel reec-
tion coecient,
0
, based on the Miller-Brown-Vegh
(MBV) model [Miller et al., 1984]:
h,v
=
0h,v
(16)
= e
2(2)
2
I
0
_
2(2)
2
(17)
=
h
w
sin
(18)
I
0
is the modied Bessel function of the rst kind,
and h
w
is the rms wave height from the Phillips ocean
wave spectrum [Phillips, 1985]:
h
w
= 0.0051v
2
w
, (19)
where v
w
is the wind speed in m/s. Within APM,
is approximated according to ITU reports [1990] by
the expression
=
1
_
3.2 2 +
_
(3.2)
2
7 + 9
, (20)
= 8
2
2
. (21)
Next is to determine the grazing angle at each PE
range step to compute the eective reection coe-
cient and subsequent impedance. Grazing angles at
the sea surface can easily be found using a geomet-
ric ray trace based on small angle approximations
to Snells law [Dockery et al., 2007]. The caveat is
that for surface-based ducting conditions, there will
be multiple grazing angles within a given range in-
terval/step, as shown in Fig. 4. Figure 4(a) shows
the refractivity prole of a 300 m surface-based duct,
and the corresponding grazing angles are shown in
Fig. 4b. Notice that beyond the skip zone, at ranges
beyond 80 km, there are multiple grazing angles (i.e.,
multiple modes) present within a given range inter-
val. The challenge is determining the proper grazing
angle associated with the dominant mode of prop-
agation at a particular range. Geometric ray trac-
ing techniques oer no further information, therefore,
spectral estimation techniques have also been used
[Dockery and Kuttler, 1996; Schmidt, 1986; Barrios,
2003] in combination with geometric ray trace meth-
ods to obtain the appropriate angle at a given range
particularly useful in complex environments where
the propagation path is a combination of sea, land,
and a range-dependent atmosphere.
Of course, one of the caveats of modeling the
impedance in this way is that for surface-based duct-
ing environments it ignores the many, equally dom-
inant, modes propagating within the duct at multi-
ple grazing angles within a range step. The advan-
tage of using the MBV method to modify the sur-
face impedance is that it is easy to implement and
for the most part has been shown to perform very
well for range-independent evaporation duct environ-
ments where the incident eld can be described, to
a very good approximation, by a single grazing an-
gle beyond the interference region [Anderson, 1995;
Rogers et al., 2000].
A more rigorous, albeit conventional, approach
has been provided by Janaswamy [2001] to model
a non-constant impedance that directly takes into
account eects of the angle-dependent reection co-
ecient present at all grazing angles. However, in
keeping with the more numerically ecient SSF PE
approach, and considering the design toward op-
erational applications, the maximum grazing angle
(shown by the dashed line in Fig. 4b) is used in com-
puting
h,v
to model rough surface eects. This re-
sults in maximum, or worst-case, clutter values and
will in general over-estimate sea clutter.
Finally, a recent approach to more accurately
model the various eld strengths at the surface, and
subsequently, clutter power described by multiple
grazing angles, has been provided by Karimian et al.
10 KARIMIAN ET AL.: RFC REVIEW
[2011b] that takes all grazing angles and their rela-
tive powers at each range-step into account.
For the RFC application, the propagation factor,
F, in the clutter equation (810) is a function of
the complex PE eld and the range (note that range
is shown by r in the clutter equations and by x in
this section, since Maxwells equations are solved in
Cartesian coordinates):
F = |u(x, z
e
)|
x, (22)
where z
e
is the eective scattering height, taken as
0.6 times the mean wave height [Reilly and Dockery,
1990], or approximately 1 m above the ocean for most
situations [Rogers et al., 2000; Barrios, 2003]. The-
oretically, F should be computed from the incident
eld at the sea surface. However, PE approximations
yield the propagation factor due to the total eld
which is close to zero at the sea surface and high fre-
quencies. Konstanzer et al. [2000] showed that the
clutter power using the total eld propagation fac-
tor at the eective scattering height is proportional
to the clutter power using the incident propagation
factor.
3.2. Sea surface reectivity models
Proper characterization of the quantity
0
F
4
in
(8) is key to providing reasonable clutter predictions
to perform RFC. The diculty is that the surface
(a)
(b)
0.1 1 10
80
70
60
50
40
30
20
S
u
r
f
a
c
e
r
e
f
l
e
c
t
i
v
i
t
y
(
d
B
)
Grazing angle (deg)
(a)
0.1 1 10
Grazing angle (deg)
(b)
GIT
TSC
SIT
BAR
Grazing angle (deg)! Grazing angle (deg)!
Figure 5. Reectivity vs. grazing angle for several sea
surface reectivity models at (a) 3 GHz, and (b) 9.3 GHz.
reectivity is implicitly dependent on the forward
propagation eects dened by F. They are inher-
ently coupled yet these two quantities are commonly
treated separately to get an estimate of the return
clutter. Most sea surface reectivity models, there-
fore, are semi-empirical and are based on site-specic
propagation data, typically with no corresponding
meteorological measurements.
There are several semi-empirical models for the av-
erage sea surface reectivity per unit area that t the
experimental sea clutter data to a function of radar
frequency, grazing angle, beam width, wind speed,
radar look direction with respect to the wind, and
polarization. This quantity, represented by
0
, is
also referred to as the normalized radar reectivity
[Nathanson et al., 1991].
A hybrid model by Barton [1988] and the Georgia
Institute of Technology (GIT) model [Horst et al.,
1978] are among the classic sea surface reectiv-
ity models for low grazing angles that are valid in
the S and X band frequencies. A comparison of
dierent models is provided in [Reilly and Dock-
ery, 1990]. GIT, Technology Services Corp. (TSC)
[Fletcher, 1978], and Barton (BAR) reectivity mod-
els at 3 GHz are compared in Fig. 5a. A similar com-
parison at 9.3 GHz with the additional Sittrop (SIT)
[Sittrop, 1977] model is shown in Fig. 5b. Notice that
the TSC, BAR, and SIT models show similar depen-
dence of
0
on grazing angle, whereas the GIT model
exhibits higher attenuation at lower grazing angles.
Lower grazing angles imply the region near the radio
horizon subject to diraction eects. The increased
attenuation shown by the GIT model as a function
of decreasing grazing angle is indicative of standard
diraction eects, and it is for this reason the GIT
model has been more widely used. That is, the GIT
reectivity can be assumed to be representative of
0
under standard atmosphere conditions.
Reilly and Dockery [1990] modied the GIT model
to consider ducting eects on the radar backscatter
by dividing
0
by the standard atmosphere propaga-
tion factor and multiplying by the propagation factor
of the desired conditions [Dockery, 1990].
Normalized mean sea backscattering coecient
0
for grazing angles of 0.1 to 60
2
are the random phase components of the complex
random variable with uniform distributions:
{log n
1
}
N
1
, G(0,
2
1
) (27)
{n
2
}
N
1
, { n
2
}
N
1
G(0,
2
2
) (28)
{
1
}
N
1
, {
2
}
N
1
, {
2
}
N
1
U(0, ) (29)
The radar output power is obtained by:
= |u
I
|
2
+|u
Q
|
2
. (30)
12 KARIMIAN ET AL.: RFC REVIEW
Thus, the observed and simulated clutter power are
related by:
o
= n
1
s
(m) +n
r
(31)
{log n
1
}
N
1
G(0,
2
1
) (32)
{n
r
}
N
1
2
(33)
where n
1
is the multiplicative noise with a lognor-
mal distribution, and n
r
is the additive receiver noise
with a
2
distribution and 2 degrees of freedom.
Working in the high CNR (clutter to noise ratio)
regime, the n
r
term can be neglected. Thus, the
modeled power in the logarithmic domain is obtained
as:
P
o
= P
s
(m) +n (34)
{n}
N
1
G(0,
2
), (35)
where, P
o
and P
s
(m) are vectors of the observed
and simulated clutter power of the prole m in dB,
and n = 10 log n
1
.
More than one source of clutter power observa-
tions can be used in an inversion. These sources can
include the clutter power at dierent frequencies, dif-
ferent radar elevation angles, or dierent snapshots
with similar conditions where P
n,o
corresponds to
the nth source of the observed clutter power. Given
N dierent sources with uncorrelated noise power
n
,
the maximum likelihood function becomes:
1
L(m) =
N
n=1
(
n
)
Nr
exp
_
P
o,n
P
s,n
(m)
2
n
_
.
(36)
Assuming that the noise power {
n
}
n=1..N
is con-
stant across dierent observations, the negative log-
likelihood function is simplied to
(m) = log L(m)
N
n=1
P
o,n
P
s,n
(m)
2
. (37)
The maximum likelihood estimate m for m is ob-
tained by minimizing (37) over the model parameter
vector m, which is similar to (23).
4.2. An inversion example
A set of refractivity prole measurements and
radar returns was recorded at Wallops Island, Vir-
ginia, April 1998 [Rogers et al., 2000; Gerstoft et al.,
2003b]. Clutter signals were measured using the
Space Range Radar (SPANDAR) with operational
frequency of 2.84 GHz, horizontal beamwidth of 0.4
,
elevation angle of 0, antenna height of 30.78 m, and
vertical polarization. The refractivity proles of the
environment were recorded using an instrumented
helicopter provided by the Johns Hopkins University
Applied Physics Laboratory. The helicopter ew in
and out along the 150
azimuth
of SPANDAR Run 07, and (b) standard atmosphere.
1
|x| = (|x
1
|, |x
2
|, ...) and x
2
=
i
|x
i
|
2
.
KARIMIAN ET AL.: RFC REVIEW 13
0 10 20 30 40 50 60
0
50
100
150
200
(a) 340 380 M
A
l
t
i
t
u
d
e
(
m
)
300 320
0
20
40
60
80
100
120
140
160
180
200
MProfile
A
l
t
i
t
u
d
e
(
m
)
(b)
Avg. heli. profile
150
inversion
Range (km)
C
l
u
t
t
e
r
p
o
w
e
r
(
d
B
)
(c)
10 20 30 40 50 60
20
10
0
10
20
30
Clutter at 150
azimuth. (b) Average of the rst 45 km of the measured prole compared to the inverted
proles of 150
.
Fig. 6 shows the propagation loss using the in-
verted prole from Fig. 7b and a standard atmo-
sphere. Surface-based ducting conditions result in
the extended range of the radar and radar fades
in unexpected locations assuming a standard atmo-
sphere. Radar parameters in this gure are identical
to those of the SPANDAR.
4.3. Bayesian approach
One important motivation behind estimation of
the refractivity structure in the environment is to
predict the radar performance in non-standard atmo-
spheric conditions. This requires the statistical prop-
erties of the parameters-of-interest such as the prop-
agation loss which can be computed from the statis-
14 KARIMIAN ET AL.: RFC REVIEW
tical properties of the atmospheric refractivity. The
unknown environmental parameters are taken as ran-
dom variables with corresponding onedimensional
(1D) probability density functions (pdfs) and an n
dimensional joint pdf. This probability function can
be dened as the probability of the model vector m
given the observed clutter power P
o
, p(m|P
o
), and it
is called the posterior pdf (PPD). The prole m with
the highest probability is referred to as the maximum
a posteriori (MAP) solution. The posterior means,
variances, and marginal probability distributions can
be found by integrating over this PPD:
i
=
_
...
_
m
i
p(m
|P
o
)dm
, (38)
2
i
=
_
...
_
m
(m
i
)
2
p(m
|P
o
)dm
, (39)
p(m
i
|P
o
) =
_
...
_
m
(m
i
m
i
)p(m
|P
o
)dm
.(40)
The posterior density of any specic environmental
parameter can be obtained by marginalizing the n
dimensional PPD as given in (7) [Kay, 1993]. Ger-
stoft et al. [2004] used importance sampling (IS)
[MacKay, 2003] to compute the necessary multi-
dimensional integrals needed to map the environmen-
tal uncertainty into propagation loss uncertainty. IS
produces unbiased distributions of the desired vari-
ables, however, the variance of the estimates depend
heavily on the importance density used in IS. An-
other problem with IS is the slow rate of convergence
for the numerical computation of the integrals. Ger-
stoft et al. [2004] also compared IS to using just the
1-D marginals of refractivity parameters to compute
the PDF of propagation loss. As long as the interpa-
rameter correlations are negligible, using marginals
is computationally more ecient than IS. They later
showed that lowering the peak clutter to noise ratio
broadens the a posteriori distribution of the propa-
gation loss [Rogers et al., 2005].
The error in IS is minimized when samples are
drawn from the posterior distribution of the en-
vironmental parameters p(m|P
o
). Sampling from
the posterior requires a Markov chain Monte Carlo
(MCMC) class sampler [
).
One of the caveats of RFC algorithms is that de-
tection of elevated ducts is not possible since the
trapped electromagnetic waves do not interact with
the sea surface. However, these ducts can be pre-
dicted based on meteorological conditions [Gossard,
1981]. The 3-D refractivity proles are intimately
linked to the weather. There have been attempts to
include climatological statistics of duct heights based
on the observation location and time of the year for
evaporation ducts [Yardim et al., 2009].
Fusion of weather prediction algorithms with RFC
inversions can greatly increase the performance of
both. An example is in costal regions when the warm
ow of air over the sea forms a rising surface duct
for radar propagation. Numerical weather prediction
(NWP) systems have undergone substantial develop-
ment in the last decade. There currently exist capa-
bilities to extract 48 h radar forecast based on output
from NWP [Marshall et al., 2008]. These forecasts
are used now to predict the radar performance [Le-
KARIMIAN ET AL.: RFC REVIEW 17
Furjah et al., 2010]. An improvement of RFC then
would be using these elds as prior into the RFC
inversion. After the inversion, the RFC posterior re-
fractivity estimates could be used to inuence the
small-scale data assimilation for NWP. More research
is required to ll the gap between weather prediction
and RFC.
Acknowledgments. This work was supported by
SPAWAR under grant number N66001-03-2-8938, TDL
0049. Authors would like to thank Dominique Lesselier
and Ted Rogers for their constructive comments.
References
Anderson, K. D., Tropospheric refractivity proles in-
ferred from low elevation angle measurements of
Global Positioning System (GPS) signals, in AGARD
Conf. Proc., pp. 2.12.7, Bremerhaven, Germany,
1994.
Anderson, K. D., Radar detection of low-altitude targets
in a maritime environment, IEEE Trans. Antennas
Propagat., 43(6), 609613, doi:10.1109/8.387177,
1995.
Babin, S. M., G. A. Young, and J. A. Carton,
A new model of the oceanic evaporation duct,
J. Appl. Meteor., 36(3), 193204, doi:10.1175/1520-
0450(1997)036<0193:ANMOTO>2.0.CO;2, 1997.
Barrios, A. E., Parabolic equation modeling in hor-
izontally inhomogeneous environments, IEEE
Trans. Antennas Propagat., 40(7), 791797,
doi:10.1109/8.155744, 1992.
Barrios, A. E., A terrain parabolic equation
model for propagation in the troposphere,
IEEE Trans. Antennas Propagat., 42(1), 9098,
doi:10.1109/8.272306, 1994.
Barrios, A. E., Advanced propagation model (APM)
computer software conguration item (CSCI), Space
and Naval Warfare Systems Center, TD 3145, 2002.
Barrios, A. E., Considerations in the development of the
advanced propagation model (APM) for U.S. Navy ap-
plications, in International Conf. on Radar, Adelaide,
Australia, 2003.
Barrios, A. E., Estimation of surface-based duct param-
eters from surface clutter using a ray trace approach,
Radio Sci., 39, RS6013, doi:10.1029/2003RS002930,
2004.
Barton, D. K., Modern Radar System Analysis, Artech
House, 1988.
Blake, L. V., Radar Range Performance Analysis, Lex-
ington Books, Lexington, MA, 1980.
Boyer, D., G. Gentry, J. Stapleton, and J. Cook, Us-
ing remote refractivity sensing to predict tropospheric
refractivity from measurements of propagation, in
Proc. AGARD SPP Symp. Remote Sensing: Valu-
able Source Inform., pp. 16.116.13, Toulouse, France,
1996.
Brooks, I. M., A. K. Goroch, and D. P.
Rogers, Observations of strong sur-
face radar ducts over the Persian Gulf,
J. Appl. Meteor., 38(9), 12931310, doi:10.1175/1520-
0450(1999)038<1293:OOSSRD>2.0.CO;2, 1999.
Budden, K. G., The Wave-Guide Mode Theory of Wave
Propagation, Prentice-Hall, Englewood Clis, NJ,
1961.
Dockery, G. D., Method for modeling sea surface clut-
ter in complicated propagation environments, in IEE
Proc. Radar and Signal Processing, vol. 137, pp. 7379,
1990.
Dockery, G. D., and J. R. Kuttler, An improved
impedance-boundary algorithm for Fourier split-
step solution of the parabolic wave equation,
IEEE Trans. Antennas Propagat., 44(12), 15921599,
doi:10.1109/8.546245, 1996.
Dockery, G. D., R. S. Awadallah, D. E. Freund, J. Z.
Gehman, and M. H. Newkirk, An overview of recent
advances for the TEMPER radar propagation model,
in IEEE Radar Conference, pp. 896905, 2007.
Dosso, S. E., and J. Dettmer, Bayesian matched-
eld geoacoustic inversion, Inverse Problems, 27(5),
055009, doi:10.1088/0266-5611/27/5/055009, 2011.
Dougherty, H., and B. A. Hart, Recent progress in duct
propagation predictions, IEEE Trans. Antennas Prop-
agat., 27(4), 542548, doi:10.1109/TAP.1979.1142130,
1979.
Douvenot, R., and V. Fabbro, On the knowledge of
radar coverage at sea using real time refractivity
from clutter, IET Radar Sonar Navig., 4(2), 293301,
doi:10.1049/ietrsn.2009.0073, 2010.
Douvenot, R., V. Fabbro, P. Gerstoft, C. Bourlier, and
J. Saillard, A duct mapping method using least square
support vector machines, Radio Sci., 53, RS6005,
doi:10.1029/2008RS003842, 2008.
Douvenot, R., V. Fabbro, P. Gerstoft, C. Bourlier, and
J. Saillard, Real time refractivity from clutter using a
best t approach improved with physical information,
Radio Sci., 45, RS1007, doi:10.1029/2009RS004137,
2010.
Doviak, R. J., and D. S. Zrnic, Doppler radar and weather
observations, p. 562, 2nd ed., Academy Press, 1993.
Fabry, F., C. Frush, I. Zawadzki, and A. Kilambi, On
the extraction of near-surface index of refraction us-
ing radar phase measurements from ground targets,
J. Atmos. Oceanic Technol., 14, 978987, 1997.
Fairall, C. W., E. F. Bradley, D. P. Rogers, J. B. Ed-
son, and G. S. Young, Bulk parametrization of air-sea
uxes for tropical ocean-global atmosphere coupled-
ocean atmosphere response experiment, J. Geophys.
Res., 101(C2), 37473764, doi:10.1029/95JC03205,
1996.
Feit, M. D., and J. A. Fleck, Light propagation
in graded-index bers, Appl. Opt., 17, 39903998,
doi:10.1364/AO.17.003990, 1978.
Fletcher, C., Clutter subroutine, Technology Service
Corp., Silver Spring, MD, USA, Memorandum TSC-
W84-01/cad, 1978.
Frederickson, P. A., K. L. Davidson, and A. K. Goroch,
Operational bulk evaporation duct model for MO-
RIAH, ver. 1.2, Tech. Rep. NPS/MR-2000-002, Nav.
Postgrad. Sch., Monterey, Calif., 2000a.
Frederickson, P. A., K. L. Davidson, C. R. Zeisse,
and C. S. Bendall, Estimating the refractive in-
dex structure parameter (C
2
n
) over the ocean using
bulk methods., J. Appl. Meteor., 39, 17701783, doi:
10.1175/1520-0450-39.10.1770, 2000b.
18 KARIMIAN ET AL.: RFC REVIEW
Geman, S., and D. Geman, Stochastic relaxation, Gibbs
distributions, and the Bayesian restoration of images,
IEEE Trans. Pattern Anal. Machine Intell., 6, 721
741, doi:10.1080/02664769300000058, 1984.
Gerstoft, P., D. F. Gingras, L. T. Rogers, and
W. S. Hodgkiss, Estimation of radio refractiv-
ity structure using matchedeld array processing,
IEEE Trans. Antennas Propagat., 48(3), 345356,
doi:10.1109/8.841895, 2000.
Gerstoft, P., L. T. Rogers, W. S. Hodgkiss, and L. J.
Wagner, Refractivity estimation using multiple el-
evation angles, IEEE J. Ocean. Eng., 28, 513525,
doi:10.1109/JOE.2003.816680, 2003a.
Gerstoft, P., L. T. Rogers, J. L. Krolik, and
W. S. Hodgkiss, Inversion for refractivity parame-
ters from radar sea clutter, Radio Sci., 38(3), 122,
doi:10.1029/2002RS002640, 2003b.
Gerstoft, P., W. S. Hodgkiss, L. T. Rogers, and
M. Jablecki, Probability distribution of low-altitude
propagation loss from radar sea clutter data, Radio
Sci., 39, RS6006, doi:10.1029/2004/RS003077, 2004.
Gingras, D. F., P. Gerstoft, and N. L. Gerr, Electromag-
netic matched eld processing: basic concepts and tro-
pospheric simulations, IEEE Trans. Antennas Propa-
gat., 42(10), 15361545, doi: 10.1109/8.633863, 1997.
Gordon, N. J., D. J. Salmond, and A. F. M. Smith, Novel
approach to nonlinear/non-Gaussian Bayesian state
estimation, IEE Proc. F, Radar and Signal Processing,
140(2), 107113, doi:10.1049/ip-f-2.1993.0015, 1993.
Gossard, E. E., Clear weather meteorological eects on
propagation at frequencies above 1 GHz, Radio Sci.,
16(5), 589608, doi:10.1029/RS016i005p00589, 1981.
Greenaert, G. L., Notes and correspondence on the
evaporation duct for inhomogeneous conditions in
coastal regions, J. Appl. Meteor. Climetol., 46(4), 538
543, 2007.
Gregers-Hansen, V., and R. Mital, An empirical sea clut-
ter model for low grazing angles, in IEEE Radar Con-
ference, pp. 15, Pasadena, CA, 2009.
Guinard, N., J. Ransone, D. Randall, C. Purves,
and P. Watkins, Propagation through an elevated
duct: Tradewinds III, IEEE Trans. Antennas Propa-
gat., 12(4), 479490, doi:10.1109/TAP.1964.1138252,
1964.
Guitton, A., and W. W. Symes, Robust inversion of seis-
mic data using the Huber norm, Geophysics, 68(4),
13101319, doi:10.1190/1.1598124, 2003.
Ha, T., W. Chung, and C. Shin, Waveform inversion us-
ing a back propagation algorithm and a Huber func-
tion norm, Geophysics, 74(3), R15R24, 2009.
Haack, T., and S. D. Burk, Summertime ma-
rine refractivity conditions along coastal Califor-
nia, J. Appl. Meteor., 40, 673687, doi:10.1175/1520-
0450(2001)040<0673:SMRCAC>2.0.CO;2, 2001.
Hardin, R. H., and F. D. Tappert, Application of the
split-step Fourier method to the numerical solution
of nonlinear and variable coecient wave equations,
SIAM Rev. 15, 423, 1973.
Helvey, R. A., Radiosonde errors and spurious surface-
based ducts, Proc. IEE F, 130(7), 643648, 1983.
Hitney, H. V., Remote sensing of refractivity structure by
direct measurements at UHF, in AGARD Conf. Proc.,
pp. 1.11.5, Cesme, Turkey, 1992.
Hitney, H. V., Refractive eects from VHF to EHF. Part
A: Propagation mechanisms, in AGARD Lecture Se-
ries, vol. LS-196, pp. 4A14A13, 1994.
Hitney, H. V., and J. H. Richter, Integrated Refrac-
tive Eects Prediction System (IREPS), Nav. Eng. J.,
88(2), 257262, 1976.
Horst, M., F. Dyer, and M. Tuley, Radar sea clutter
model, in Int. IEEE AP/S URSI Symposium, vol.
Proc. Part 2 (A80-15812 04-32), pp. 610, London In-
stitute of electrical engineers, London, England, 1978.
Hua, Y., and W. Liu, Generalized Karhunen-Loeve trans-
form, IEEE Signal Process. Lett., 5(6), 141142, 1998.
Huber, P. J., Robust regression: Asymptotics, conjec-
tures, and Monte Carlo, Ann. Statist., 1(5), 799821,
1973.
ITU reports, Report 1008-1, Reection from the surface
of the Earth, in Recommendations and reports of the
CCIR, Propagation in nonionized media, vol. 5, 1990.
Janaswamy, R., Radio wave propagation over a non-
constant immittance plane, Radio Sci., 36, 387405,
doi:10.1029/2000RS002338, 2001.
Jensen, F. B., W. A. Kuperman, M. B. Porter, and
H. Schmidt, Computational ocean acoustics, Springer
London, Limited, 2011.
Jeske, H., State and limits of prediction methods for
radar wave propagation conditions over the sea, in
Modern topics in Microwave Propagation and Air-Sea
Interaction, edited by A. Zancla, pp. 130148, D. Rei-
del Publishing, 1973.
Julier, S., J. Uhlmann, and H. F. Durrant-White,
A new method for nonlinear transformation of
means and covariances in lters and estima-
tors, IEEE Trans. Automat. Control, 45, 477482,
doi:10.1109/9.847726, 2000.
Karimian, A., C. Yardim, P. Gerstoft, W. S. Hodgkiss,
and A. E. Barrios, Estimation of refractivity using a
multiple angle clutter model, Radio Sci., submitted,
2011a.
Karimian, A., C. Yardim, P. Gerstoft, W. S. Hodgkiss,
and A. E. Barrios, Multiple grazing angle sea clutter
modeling, IEEE Trans. Antennas Propagat., submit-
ted, 2011b.
Katzin, M., R. W. Bauchman, and W. Binnian, 3- and 9-
centimeter propagation in low ocean ducts, Proc. IRE,
35, 891905, 1947.
Kay, S. M., Fundamentals of statistical signal processing,
volume I: Estimation theory, Prentice Hall, Englewood
Clis, NJ, 1993.
Kerr, D. E., Propagation of short radio waves, McGraw-
Hill, 1951.
Ko, H. W., J. W. Sari, and J. P. Skura, Anomalous
microwave propagation through atmospheric ducts,
Johns Hopkins APL Tech. Dig. 4(2), 1226, 1983.
Konstanzer, G. C., J. Z. Gehman, M. H. Newkirk, and
G. D. Dockery, Calculation of the surface incident eld
using TEMPER for land and sea clutter modeling,
Proc. on Low Grazing Angle Clutter: Its Character-
ization, Measurement and Application, RTO-MP-60,
pp. 141142, 2000.
KARIMIAN ET AL.: RFC REVIEW 19
Kraut, S., R. Anderson, and J. Krolik, A gener-
alized Karhunen-Loeve basis for ecient esti-
mation of tropospheric refractivity using radar
clutter, IEEE Trans. Sig. Proc., 52(1), 4859,
doi:10.1109/TSP.2003.820297, 2004.
Krolik, J. L., and J. Tabrikian, Tropospheric refractivity
estimation using radar clutter from the sea surface, in
Proc. Battlespace Atmospheric Conference, SSC-SD,
pp. 635642, San Diego, CA, 1997.
Kukushkin, A., Radio wave propagation in the marine
boundary layer, Wiley, 2004.
Kuttler, J. R., Dierences between the narrow-angle
and wide-angle propagators in the split-step Fourier
solution of the parabolic wave equation, IEEE
Trans. Antennas Propagat., 47(7), 11311140, doi:
10.1109/8.785743, 1999.
Kuttler, J. R., and R. Janaswamy, Improved Fourier
transform methods for solving the parabolic wave
equation, Radio Sci., 37(2), 10211031, doi:
10.1029/2001RS002488, 2002.
LeFurjah, G., R. Marshall, T. Casey, T. Haack, and
D. De Forest Boyer, Synthesis of mesoscale numeri-
cal weather prediction and empirical site-specic radar
clutter models, IET Radar Sonar Navig., 4(6), 747
754, 2010.
Leontovich, M. A., and V. A. Fock, Solution of the prob-
lem of electromagnetic wave propagation along the
earths surface by the method of parabolic equation,
Acad. Sci. USSR J. Phys., 10, 1323, 1946.
Levy, M., Parabolic equation methods for electromagnetic
wave propagation, The Institution of Electrical Engi-
neers, London, 2000.
Lin, L., Z. Zhao, Y. Zhang, and Q. Zhu, Tropo-
spheric refractivity proling based on refractivity pro-
le model using single ground-based global position-
ing system, Radar Sonar Navigation, IET, 5(1), 711,
doi:10.1049/iet-rsn.2009.0167, 2011.
Liu, W. T., K. B. Kataros, and J. A. Businger, Bulk pa-
rameterization of air-sea exchanges of heat and water
vapor including the molecular constraints at the inter-
face, J. Atmos. Sci., 36, 17221735, doi:10.1175/1520-
0469(1979)036<1722:BPOASE>2.0.CO;2, 1979.
Lowry, A. R., C. Rocken, S. V. Sokolovsky, and K. D.
Anderson, Vertical proling of atmospheric refractiv-
ity from ground-based GPS, Radio Sci., 37, 10411059,
doi:10.1029/2000RS002565, 2002.
MacKay, D. J. C., Information Theory, Inference and
Learning Algorithms, Cambridge University Press,
Cambridge, United Kingdom, 2003.
Marshall, R. E., W. D. Thornton, G. LeFurjah, and
S. Casey, Modeling and simulation of notional future
radar in non-standard propagation environments fa-
cilitated by mesoscale numerical weather prediction
modeling, Naval Engineers Journal, 120(4), 5566,
doi:10.1111/j.1559-3584.2008.00165.x, 2008.
Mentes, S., and Z. Kaymaz, Investigation
of surface duct conditions over Istanbul,
Turkey, J. Appl. Meteor. Climetol., 46, 318337,
doi:10.1175/JAM2452.1, 2007.
Metropolis, N., A. W. Rosenbluth, M. N. Rosenbluth,
A. H. Teller, and E. Teller, Equation of state cal-
culations by fast computing machines, J. of Chemical
Physics, 21, 10871092, doi:10.1063/1.1699114, 1953.
Miller, A. R., R. M. Brown, and E. Vegh, New deriva-
tion for the rough surface reection coecient and for
the distribution of the sea-wave elevations, Proc. Inst.
Elect. Eng., 131(2, pt.H), 114116, doi:10.1049/ip-h-
1.1984.0022, 1984.
Miller, E. L., and A. S. Willsky, A multiscale, statistically
based inversion scheme for linearized inverse scattering
problems, IEEE Trans. Geosci. Remote Sens., 34(2),
346357, doi:10.1109/36.485112, 1996a.
Miller, E. L., and A. S. Willsky, Wavelet-based meth-
ods for the nonlinear inverse scattering problem using
the extended Born approximation, Radio Sci., 31(1),
5165, doi:10.1029/95RS03130, 1996b.
Nathanson, F. E., J. P. Reilly, and M. N. Cohen, Radar
design principles: Signal processing and the environ-
ment, 2 ed., McGraw-Hill, 1991.