You are on page 1of 22

International Journal of Remote Sensing

Vol. 26, No. 21, 10 November 2005, 47634784

Statistical estimate of the hourly near-surface air humidity in


eastern Canada in merging NOAA/AVHRR and GOES/IMAGER
observations
K.-S. HAN*{, A. A. VIAU{, Y.-S. KIM{ and J.-L. ROUJEAN
{Department of Satellite Information Science, Pukyong National University, 599-1
Daiyeon-3 Dong, Nam Gu, 608-737 Busan, Korea
{Departement des Sciences Geomatiques, Laboratoire de Geomatique Agricole et
dAgriculture de Precision, Pavillon Louis-Jaques Casault, Universite Laval,
Quebec (Quebec), G1K 7P4, Canada
METEO-FRANCE, CNRM/GMME/MATIS, 42 Avenue Gaspard Coriolis, 31057
Toulouse Cedex, France

(Received 13 September 2004; in final form 21 March 2005 )

Estimates of the relative humidity near the ground are frequently requested by
scientific communities concerned about weather forecasting, disease prediction,
and agriculture. To face the dearth of meteorological observations provided by
synoptic networks, remote sensing measurements are particularly useful,
specifically because they can provide coherent information at a regional
representative scale. The present investigation gives an update on the potential
for using satellite data to estimate the near-surface relative humidity. The
IMAGER sensor on board the Geostationary Operational Environmental
Satellite (GOES) is used to obtain the hourly infrared datasets. In addition,
data from the Advanced Very High Resolution Radiometer (AVHRR) flown on
the National Oceanic and Atmospheric Administration (NOAA) Sun-synchronous satellite series is used to calculate the daily normalized difference vegetation
index (NDVI). Estimates of the relative humidity are assessed using various
variables like the surface temperature, NDVI, the precipitable water, the digital
elevation model, the date and local time. The study approach combines
empirically these variables into third-order polynomial multiple regressions with
stepwise functions. The data are split in two parts: the algorithm development
dataset and the validation dataset. The estimation model is developed by a
stepwise function, which selects independent variables and decides corresponding
coefficients. The model validity is further assessed by employing a comparison
with the results obtained from the model output using a validation dataset. The
accuracy achieved using the validation dataset is in a good agreement with
development dataset accuracies. The relative humidity accuracy derived from the
present method is within 10% compared to field measurements. The largest
discrepancies between model and measurements were observed over forested
areas. One outcome from this study is that the difference in results between
forested and non-forested targets is enhanced with the topography.

*Corresponding author. Email: kyungsoohan@yahoo.com


International Journal of Remote Sensing
ISSN 0143-1161 print/ISSN 1366-5901 online # 2005 Taylor & Francis
http://www.tandf.co.uk/journals
DOI: 10.1080/01431160500177711

4764
1.

K.-S. Han et al.

Introduction

The relative humidity measured at meteorological stations is a key descriptor of


near-surface atmospheric and environmental conditions (Geiger 1965, Rosenberg
et al. 1983, Monteith and Unsworth 1990). Routine estimates of relative humidity
are particularly relevant for applications in domains like weather forecasting,
disease prediction, and agricultural science. The analysis of spatial and temporal
patterns of relative humidity variability near the ground leads to a better
understanding of the processes linking climate issues and terrestrial life. Actually,
the distribution of the near-surface relative humidity is closely related to dynamics
patterns of surface properties that vary in time and space.
Relative humidity is physically defined as the ratio between the actual mixing
ratio and the saturation mixing ratio (Hess 1959). It is currently derived from the air
and dew point temperatures measured at synoptic meteorological networks.
However, the utility of these measurements may be questionable for appropriately
representing the spatial detail in processes occurring over a large region such as the
one studied here. In fact, meteorological station density is low in many parts of
the world. Furthermore, real time access to the relative humidity fields may be
hampered by both technical (no acquisition, subsequent delay in dissemination) and
cost issues (Jones 1995, Karl et al. 1995). The assessment of hydrological variables
from satellite measurements, such as evapotranspiration, relies on a kriging
interpolation of relative humidity fields, not applicable for pixel-by-pixel
implementation. The reliability of this kriging interpolation is normally constrained
by the sparseness of our measurement networks. In this regard, estimates of relative
humidity adequately derived from remote sensing offer distinct advantages for
thorough hydrometeorological studies.
During the last decade, several studies have focused on efficient methods for
producing routine estimates of air temperature from surface temperature that take
advantage of remote sensing sensor systems devoted to Earth observation. Vogt
et al. (1997) used a statistical relationship between the surface temperature sensed by
satellite and the daily air temperature measured at ground stations to improve the
regional survey of air temperature measurements. Prihodko and Goward (1997)
attempted to estimate the air temperature based on the relationship that exists
between the skin surface temperature and the normalized difference vegetation index
(NDVI). Both are derived from the National Oceanic and Atmospheric
Administration/Advanced Very High Resolution Radiometer (NOAA/AVHRR)
observing system. This latter method, however, is limited by the fact that the
relationship at night is inverted as compared to daytime. Other studies successfully
estimated the maximum air temperature using multiple statistical analyses from a
digital elevation model (DEM) into which inputs are the digital land use and the
surface temperature (Viau et al. 1996, 2000, Paquet et al. 1997).
Relative humidity, when combined with air temperature, can be used to estimate
the actual amount of moisture in the atmosphere, a quantity sometimes referred to
as precipitable water. The geographic distribution of precipitable water is important
because water vapour acts as a greenhouse gas absorbing long-wave infrared
radiation emitted from the Earth system. In desert areas, the air is mostly dry and
cannot absorb the outgoing heat. This partly explains why desert temperatures are
much lower at night-time than for other regions. Several authors (Reitan 1963,
Smith 1966) found a correlation between the mean monthly total precipitable water
and the mean monthly surface dew point temperature at meteorological observation

Estimating near-surface air humidity in eastern Canada

4765

stations. Choudhury (1996) proposed coefficients for a formula, reported by Reitan


(1963), from a monthly database of measured precipitable water and dew point
temperature for a site in Canada. Recent studies (Prince and Goward 1995, Prince
et al. 1998) used Smiths method (1966) to determine the vapour pressure deficit
from precipitable water values retrieved by satellite sensors. In spite of these efforts,
the applicability of this algorithm developed for monthly estimates remains
questionable for daily or hourly retrievals. In contrast to air temperature studies,
investigations that seek to derive an estimate of the near-surface relative humidity
from satellites are rare. Two factors make remote sensing of relative humidity
particularly difficult. First, inferring near-surface relative humidity directly from
remotely sensed data is difficult in that the vertical resolution of water-vapour
sounding instrument is not sufficient (Aumann et al. 2003). Second, investigations
have not found remotely sensed parameters that linearly match spatial and temporal
variations in relative humidity. Therefore, a statistical approach using multiple
parameters to measure relative humidity from satellite-derived data seems highly
desirable.
In this context, the present study has three main objectives: to identify the
remotely sensed data that contribute most strongly to accurate estimates of relative
humidity; to develop a semi-empirical model using a multiple regression polynomial
analysis that selects statistically significant satellite variables linked to near-surface
humidity; and to estimate the relative humidity at hourly temporal resolution based
on a regression. High temporal resolution estimates could be possible using
geostationary satellite data such as those provided by the Geostationary
Operational Environmental Satellite (GOES). In addition, the NOAA Polar
Operational Environmental Satellite (NPOES) provides visible and near-infrared
data on a daily basis that can be used to advantage in obtaining parameters such as
the vegetation state or condition using the NDVI. Sections 2 and 3 detail datasets
used in the present study, satellite-derived parameters, and the background of the
algorithm. Section 4 presents the estimation model established using the development dataset whereas 5 yields a comparison with the estimation resulting from the
validation dataset. Section 6 concludes the study, discusses the limitations, and
proposes a future direction.
2.
2.1

Material and parametrization


Dataset and image processing

The study area is located in the western part of Quebec province in Canada. It is
delimited in longitude by 71.47u and 75.65u west coordinates and in longitude by
45.03u and 49.74u north coordinates. In this region, dominant landscape units
encompass forest and agriculture areas. Adequate datasets were acquired for the
purpose of this study and used for the development and validation of the retrieval
algorithms. The data used for this research work were collected during two periods,
110 June and 110 July 1997, thereby covering the beginning and middle of the
summer growing season. Meteorological data included hourly air temperature (Ta),
dew point temperature (Td), and relative humidity (RH) observations from 41
automatic surface stations. The set of stations belongs to three different regional
networks: (1) Environment Canada, (2) the POMMIER network of MAPAQ
(Ministe`re de Agriculture, des Pecheries et de lAlimentation du Quebec), and (3)
SOPFEU (Societe de Protection des Forets contre le Feu). Environment Canada,

4766

K.-S. Han et al.

Figure 1. Distribution of the meteorological stations of three different regional networks in


the study area(1) Environment Canada, (2) the POMMIER network of MAPAQ
(Ministe`re de Agriculture, des Pecheries et de lAlimentation du Quebec), and (3) SOPFEU
(Societe de Protection des Forets contre le Feu).

POMMIER, and SOPFEU contain 11, 9, and 21 stations, respectively, in the region
of interest (figure 1). All measuring instruments for three networks are configured
according to Environment Canada and World Meteorological Organization
(WMO) regulations.
The present study also incorporates satellite data from the AVHRR sensor on
board the NOAA-12 and NOAA-14 Sun-synchronous satellites and from the
IMAGER sensor flown on the geosynchronous GOES-8 satellite. All data were
obtained from the Atmospheric Environment Service of Environment Canada. To
remove cloud effects from the signal, we used the five channels of NOAA-12 and

Estimating near-surface air humidity in eastern Canada

4767

Table 1. Systemic and optical characteristics of the GOES/IMAGER and NOAA/AVHRR


sensors.
Satellite
Sensor
orbit type
altitude
repetitiveness
Channel
1
2
3
4
5

GOES-8

NOAA-12, -14

IMAGER
geosynchronous
<35 788 km
30 min

AVHRR
Sun-synchronous
<833 km
<12 h

Wavelength
(mm)
VIS
IR
IR
IR
IR

0.550.75
3.84.0
6.57.0
10.211.2
11.512.5

Resolution at
nadir (km2)

Wavelength
(mm)

Resolution at
nadir (km2)

1.061.0
4.068.0
4.068.0
4.068.0
4.068.0

VIS
0.580.68
NIR 0.7251.10
IR
3.553.93
IR
10.311.3
IR
11.512.5

1.161.1
1.161.1
1.161.1
1.161.1
1.161.1

VIS: visible channel, NIR: near-infrared channel, IR: infrared channel.

-14 and channels 1, 2, 4 and 5 of GOES-8. Sensor characteristics are summarized in


table 1. These satellite data have a spatial resolution about 1 km by 1 km at nadir for
AVHRR and IMAGER channel 1, and 4 km by 4 km for the three other IMAGER
channels (table 1). The study period included 913 IMAGER scenes that provided
4116 total images in different channels and 52 AVHRR scenes that provided 260
total images in different channels. The NOAA satellite series provided between two
and three AVHRR scenes per day at irregular time intervals. Visible and nearinfrared AVHRR channels were used to derive the NDVI images which are
composed on a daily basis. IMAGER channels 4 and 5 are available twice per hour.
This is necessary to measure precipitable water (Pw) and surface temperature (Ts).
AVHRR thermal infrared channels can also provide Pw and Ts, but GOES/
IMAGER data were used because of their finer temporal resolution. Finally, a
DEM was created from Canadian topographic maps of Geomatics Canada with a
scale of 1 : 25 000 and 30-m contour intervals.
Digital signals in each channel were converted to an apparent radiance. Planetary
reflectance or brightness temperatures were then calculated for each individual
channel using the calibration coefficients. Radiance calibration coefficients were
taken from Weinreb et al. (1997) for GOES and from Kidwell (1998) for NOAA.
Calculations included illumination and view-angle effects (Teillet 1992). Geometric
correction was further applied to the whole imagery with the PCI Geomatica 8.0
(PCI Geomatics Enteprises Inc.) image processing software based on the nearestneighbour resampling method. The correction has r.m.s. errors (RMSE) of 0.74 km
and 0.60 km for GOES/IMAGER and NOAA/AVHRR, respectively, in both the xand y-directions. An important step of the data-processing stream concerns cloud
detection, and removal of residual cloud contamination. Most cloud-detection
algorithms use successive tests with different channels in the visible and infrared
parts of the spectrum (Caselles et al. 1998). Cloud masking in the present study
involves use of a threshold method based on simplified versions of several
algorithms (Saunders 1986, Gutman et al. 1987, Saunders and Kriebel 1988, Shin
et al. 1996). Correction for atmospheric effects was accomplished by using the
radiative transfer model MODTRAN 3.7 (Acharya et al. 1998). For model
operation, use of a standard atmospheric profile is necessary when the radiosounding measurements are not available. MODTRAN 3.7 provides several

4768

K.-S. Han et al.

standard atmospheres having different profiles. This standardization of the


atmosphere could introduce some error in the results because: the standard
atmospheres are truncated at 10 km and no absorption is assumed above this level;
there is a difference between the standard atmospheric profile and the one that exists
at the time measurement. Nevertheless, previous investigations (Richter 1990,
Sobrino et al. 1991, Teillet 1992, Perry and Moran 1994) have demonstrated the
relevance of standard atmospheres to perform atmospheric corrections. In
particular, Richter (1990) presented a maximum absolute error of 0.03 for
Thematic Mapper (TM) due to model approximations as a function of surface
reflectances using standard atmosphere for mid-latitude summer. According to these
results, it was considered acceptable to use the model atmosphere in MODTRAN
for application to this study. Considering the study area and period, a standard
atmosphere of mid-latitude summer category was selected to get the atmospheric
profile of pressure, temperature, air density, H2O, O3, N2O, CO, and CH4. A
corrected ground temperature was derived from the measured radiance using
Plancks law according to the blackbody theory.
2.2

Derivation of the physical variables

The split-window algorithm proposed by Ulivieri et al. (1994) for the AVHRR was
used to retrieve Ts. This combines brightness temperatures in two spectrally adjacent
channels. It is worth outlining that the atmospheric absorption and the surface
emissivity are inherently accounted for in the algorithm. The algorithm reads:
Ts ~T4 z1:8T4 {T5 z481{e{75De

where T4 and T5 are the brightness temperatures of channels 4 and 5, respectively,


and e and De are the mean and difference for AVHRR channels 4 (10.5 mm) and 5
(11.5 mm) surface emissivities.
IMAGER thermal infrared channels 4 and 5 have similar wavelengths to the
AVHRR as shown in table 1 (GOES-12 IMAGER does not have channel 5). Han
et al. (2004) showed that Ts computed with IMAGER data agreed well with Ts
computed with AVHRR data, using the algorithm developed by Ulivieri et al.
(1994). A RMSE of 0.9 K was found between estimations using both satellites
datasets. We therefore assumed that the split-window algorithms developed for
AVHRR are valid for IMAGER data as well. The split-window algorithm of
Choudhury et al. (1995) is used, therefore, for estimating Pw over land, which as
expressed is
Pw~

T4 {T5 {a
b

2a

where
a~75:De0 z8:1{e4 {0:15

2b

b~0:05:1{44:De0 {5:1{e4 

2c

De0 ~e4 {e5 =e5

2d

Channel emissivities in this split-window algorithm were calculated using the


empirical method proposed by Cihlar et al. (1997). They proposed a loglinear

Estimating near-surface air humidity in eastern Canada

4769

relationship between NDVI, e4, and De:


e4 ~0:9897z0:029:lnNDVI

De~0:01019z0:01344:lnNDVI

NDVI and emissivity are assumed to be constant during the day for any pixel.
However, in rangelands there is an increase in photosynthetic activity, or
greenness, after rainfall. Dew or rainfall event could change somewhat the
emissivity. Therefore, maximum NDVI values composed within a day for all
daytime NDVI images were used in the above emissivity calculation. To match
AVHRR and IMAGER spatial resolutions for the purpose of producing reliable
estimates of Ts and Pw, 16 AVHRR emissivity pixels were averaged and further
integrated into one single pixel. GOES data were available 15 and 45 min after
hourly acquisition. Ts and Pw images at 15 min interval before and after an hourly
acquisition were averaged to provide an hourly database (i.e. 00 h, 01 h, 02 h, ,
23 h).
3.

Algorithm overview for relative humidity estimation

RH is defined as the ratio of the mixing ratio to the saturation mixing ratio (Hess
1959). Assuming water vapour is an ideal gas, this definition leads to:
RH~ea =e0

5a

ea ~0:06108exp17:27:Td =273:3zTd 

5b

e0 ~0:06108exp17:27:Ta =273:3zTa 

5c

where

where ea is the vapour pressure of air (kPa), e0 is the saturation vapour pressure
(kPa), Ta is the air temperature (uC) and Td is the dew point temperature (uC).
Therefore, RH is a function of Ta and Td. Although Ta and Td can be estimated from
satellite sensor data, the subsequent calculation of RH has not been performed. The
RH estimation through independent retrievals of Ta and Td may significantly
decrease the accuracy because of the error amplified by double error sources (Ta and
Td estimation procedures). To minimize and avoid these errors, the algorithm
developed in this paper directly estimates RH. This study attempts to provide an
hourly database and to derive an equation that considers both temporal and spatial
information. The algorithm presented here is designed with various input variables
identified using a multiple regression approach.
Too many independent variables (or predictors) may be addressed when various
input variables are used in a multiple regression approach with high-order
polynomial. Therefore, a mean to reduce the number of predictors is strongly
requested. The stepwise function analysis, therefore, was employed for selecting
predictors. Draper and Smith (1981) developed a stepwise multiple regression for a
multivariate dataset. The stepwise polynomial regression function is an alternative
to prediction of mining functions. It is very efficient, even when applied to large
amounts of data. Stepwise multiple regression is also advantageous because model
optimizations can be applied. Such methods discard insignificant variables so that
they have no influence on the prediction model. It is, therefore, an enhanced

4770

K.-S. Han et al.

regression method based on multiple polynomial regression. Multiple


polynomial regression predicts the value of a dependent variable on the basis of
n independent variables, each of which is expressed as a polynomial of the
mth degree. The simplified model below shows the general model without the tied
terms:
X
6
y~Az
f Xi ze
where f Xi ~Bi1 Xi1 zBi2 Xi2 z . . . :zBim Xim , y is the dependent variable, Xi are the
independent variables, A and Bi are the unknown coefficients, and e is the error
term.
The stepwise polynomial regression function retrieves the most significant
variables. The f(Xi) of the fitting model is
2
3
Bi1 Xi1
6
7
Bi1 Xi1 zBi2 Xi2
6
7
f Xi ~6
7
7
4
5
:
Bi1 Xi1 zBi2 Xi2 z . . . :zBim Xim
The number of available terms is a function of the degree of the polynomial. The
method allows a significance threshold so that independent variables with
probability values above a threshold will not be considered when the regression
model is built. Thus, only those independent variables that contribute most
significantly to the result are incorporated into the model. This regression function
first selects the most significant independent variable and builds an initial regression
model. The function then checks whether this selection diminished the significance
of other variables. If this is the case, the variable is removed from the model. After
that, the function selects the independent variable with the second highest
significance. This process is repeated until all variables whose significance is equal
to or lower than the specified threshold are discarded or added to the model (http://
www-3.ibm.com/software/data/iminer/fordata/). In the final step, an estimation
equation should be selected from all derived models as equation (7) by the end-user.
To find the input variables used in stepwise function, all relationships which exist
among satellite parameters and static variables are considered and fully described in
the next section.
Before applying the above-described algorithm, the data were divided into
two parts. The first part includes data from 110 June 1997 and was used for
algorithm development. The second part includes data from 110 July 1997 and was
used for model validation. The development dataset included 4283 cloud-free
observations and the validation dataset included 2761 cloud-free observations.
Figure 2 outlines the methodology used in the present study, which can be explained
as follows:
(1)

(2)
(3)

selection of input variables and optimum number of degree of freedom for


the polynomial fit (based on previous investigations and/or various
relationships);
selection of estimated variables based on a stepwise regression analysis and
derivation of output coefficients;
estimation of RH with the final independent variables and the corresponding coefficients (the estimation accuracy was assessed);

Estimating near-surface air humidity in eastern Canada

4771

Figure 2. Architectural design of the method for establishing the estimation model and for
its validation.

(4)

(5)

4.
4.1

test of established RH estimation model derived from the development


dataset and applied to the validation dataset (the same coefficients should be
used in the validation procedure); and
assessment of validation accuracy (the accuracy of development and
validation datasets was compared).

Analysis
Physical discussion for input parameter selection

4.1.1 Ta, Td and RH. Dew points indicate the amount of moisture in the air. The
higher the dew points, the higher the moisture content of the air at a given
temperature. A state of saturation exists when the air is holding the maximum
amount of water vapour possible at the existing temperature and pressure. When Td
and Ta are equal, the air is said to be saturated. Td is never greater than the Ta.
Therefore, if the air cools, moisture must be removed from the air and this is
accomplished through condensation. RH can be inferred from dew point values. The
quantity RH is calculated from a known Ta and Td. RH was high (low) when the
difference between Ta and Td was small (large) (figure 3(a)). When Ta and Td are
very close, the air has a high RH (around 16 oclock in figure 3(a)). The opposite is
true when there is a large difference between Ta and Td, which indicates air with
lower RH (around 1 oclock in figure 3(a)). Locations with high RH indicate that the
air is nearly saturated with moisture; clouds and precipitation are therefore quite
possible. Weather conditions at locations with high Td are likely to be uncomfortably humid. If we can accurately estimate Ta and Td, it is easy to obtain RH.
However, as mentioned in the previous section, independent retrievals of Ta and Td
may significantly amplify the error in RH estimation by double error sources from
Ta and Td estimation procedures. Therefore, it is necessary to construct a direct

4772

K.-S. Han et al.

Figure 3. Relationships between parameters for 2 June 1997: (a) relationship between Ta,
RH, and Td on a daily scale; (b) relationship between Td and Pw; (c) relationship between Ta
and Ts; (d ) relationship between Ta and DEM; (e) (Ts2Ta) variation according to daily
NDVI; and ( f ) diurnal variation of Ta and Ts at LAcadie site.

estimation RH algorithm that employs ensemble the parameters related to Ta and/or


T d.
4.1.2 Relationship between Td and Pw. Pw is defined as the amount of water in a
vertical column of atmosphere. The unit of measure is typically the depth to which
the water would fill the vertical column if it were condensed to a liquid. For
example, high Pw (in the absence of clouds) indicates a very moist atmosphere.
Reitan (1963) firstly found a logarithmic relationship between Pw and Td for
monthly estimation. For this relationship, Choudhury (1996) presented the
coefficients optimized for the study area. Although they developed this algorithm
for just the monthly time scale, these studies offered an idea that the water amount
of vertical column could impact someway on humidity near the surface. We

Estimating near-surface air humidity in eastern Canada

4773

obtained a linear correlation (determination coefficient (R2)50.54) between hourly


values of Pw and Td (figure 3(b)). In hourly scale, these two parameters do not show
the strong relationship which resulted for monthly estimation. However, from
figure 3(b), we can consider that Pw is one of the factors which affect Td, even if it is
not the absolute major factor.
4.1.3 Influential factors on Ta. The Ts is an estimate based upon the long-wave
radiation emitted from the Earths surface, and is not a true indication of air
temperature. The infrared signal detected by satellite is effectively a mixed signal,
and is the sum of different fractions of energy. These fractions include the energy reemitted from the atmosphere, as well as energy emitted from the ground. In
addition, during the day, a large proportion of the energy detected by satellite will be
attributable to re-emitted long-wave radiation, resulting from solar insolation. In
addition, some energy will be absorbed by the atmosphere as it travels out towards
space (Cresswell et al. 1999). Several studies (Caselles et al. 1991, Martin 1997, Vogt
et al. 1997, Viau et al. 2000) have used a simple method to improve the representativeness of daily maximum air temperature measurements at regional scales using
only satellite-derived Ts. To improve the accuracy of the linear model using a
relationship with Ts, Paquet et al. (1997) proposed a multiple regression model to
estimate daily Ta using Ts, DEM, DEM-derived slope and aspect. In this study, it
was difficult to find a variable with a simple relationship with hourly Ta. A linear
relationship between hourly Ta and Ts values resulted for all stations and study
periods (figure 3(c)). Although Ta can be derived from Ts within a standard error of
estimation (SEE) of about 3uC (figure 3(c)), the relationship with Ts was not
sufficient to estimate Ta because of broadly distributed residuals, i.e. large
uncertainties.
The effect of elevation should be considered. The relation between Ta and
DEM reflects the fact that Ta cools at 0.6uC/100 m in the lower atmosphere
(the moist adiabatic lapse rate). Figure 3(d ) shows the distribution of Ta as a
function of elevation for all stations in the study. The regression line represents
a lapse rate of 0.57uC/100 m, which is very similar to the moist adiabatic lapse
rate.
Another option is the temperature/vegetation index (TVX) concept based on
empirical observations of a strong negative correlation between Ts and NDVI
(Goward et al. 1985, Goward and Hope 1989, Nemani and Running 1989, Price
1990, Smith and Choudhury 1991, Hope and McDowell 1992, Nemani et al. 1993).
According to TVX air temperature hypothesis presented by Prihodko and Goward
(1997), Ts for an infinitely dense green canopy (maximum NDVI) evaluated from
the Ts/NDVI plot should approximate the air temperature (Czajkowski et al. 1997).
That is, Ts and Ta are very close for dense vegetation canopies with a high NDVI
value. Figure 3(e) shows the variation of (Ts2Ta) at noon time versus the daily
NDVI. For NDVI values less than 0.15, the quantity (Ts2Ta) ranges between 0 and
22uC whereas it remains between 0 and 5uC for NDVI values beyond 0.5. When
daily NDVI is increasing, (Ts2Ta) is decreasing.
Finally, it is necessary to take into account the diurnal cycle of Ta and Ts. During
the daytime, surface temperature exceeds the air temperature, while the air
temperature exceeds the surface temperature at night (figure 3( f )). Figure 3( f )
shows the diurnal cycle of Ta and Ts in LAcadie (45.3u N, 73.36u W) on 2 June 1997.
A third-order polynomial is required to mimic the diurnal variations of Ta and Ts
with R250.91 and 0.93 for Ts and Ta, respectively.

4774
4.2

K.-S. Han et al.


Model development

Based on the aforementioned relationships, hourly RH is estimated here with six


input parameters: Ts, Pw, NDVI, DEM, local time (LT), and Julian day (JD). In
further developments, we consider 4283 cloud-free observations. Third-order
combinations of the six input variables yielded 84 independent variables. There
were 44 different statistical models that had a significance threshold of 0.05. Figure 4
shows statistical results for the various models. The best results are obtained with
the model 44 (RMSE510.49% and R250.77) with 24 independent variables.
However, model 44 slightly improves the scores with model 13 (RMSE510.8% and
R250.73) with 11 independent variables (figure 4). Nevertheless, model 44 is
adopted because the best significance level is achieved. Independent variables
selected using the stepwise or best-subset multiple regression analysis and their
significance levels are shown in table 2. All significance levels of the chosen model
are below 0.004. The selected estimation model using 24 final independent variables
can be written as:
RHe ~b0 z

24
X

bi Fi Ts , NDVI, Pw, DEM, LT, JD

i~1

where RHe is the estimated RH, Fi are output variables combining input variables,
b0 is a constant, and bi is the regression coefficient of the ith independent variable.
The RHe values estimated from equation (8) agreed with measured values; the
regression line slope was 0.998; the intercept was 0.114%; R2 was 0.77; and the SEE
was 10.49%.

Figure 4. Comparison of models detected by stepwise method for R2 and the number of
independent variables. The dotted lines with number labelling indicate the scale for the bar
graph (in grey) of number of variables.

Estimating near-surface air humidity in eastern Canada

4775

Table 2. Statistics from the stepwise function for the development dataset.
Independent
variable (Fi)

Significance level

Unstandardized
coefficient (bi)

Variation of bi

Standardized
coefficient

Constant
F15x1
F25x1x2x3
F35x1x32
F45x1x3x5
F55x1x3x6
F65x1x4x6
F75x33
F85x22
F95x23
F105x2x3x4
F115x2x3x6
F125x3
F135x32
F145x32x2
F155x32x6
F165x3x4x6
F175x4
F185x42x1
F195x42x2
F205x43
F215x42x6
F225x62x2
F235x62x3
F245x62x4

<0
<0
8.965610204
1.221610207
1.224610223
1.162610207
<0
<0
3.130610212
5.792610213
4.209610203
4.078610205
<0
<0
<0
2.289610204
1.679610210
<0
<0
1.238610208
<0
3.728610212
7.675610205
2.626610207
1.715610213

2101.7560
0.0791
0.0007
0.0001
20.0005
20.0001
0.0076
0.0002
4.5929
25.6415
20.0953
20.0042
0.1149
20.0098
0.0036
0.0001
20.0134
349.7758
20.6921
25.9881
2605.3060
0.4812
0.0019
0.0002
20.0121

5.8740
0.0080
0.0002
1.08610205
4.52610205
2.08610205
0.0010
1.15610205
0.6880
0.8120
0.0330
0.0010
0.0090
0.0004
0.0004
2.75610205
0.0020
20.5920
0.0480
1.0490
37.5700
0.0720
0.0005
4.36610205
0.0020

3.171
0.178
0.784
20.101
21.442
2.633
3.665
1.499
20.938
20.479
20.697
3.543
27.166
0.924
0.982
22.649
10.115
25.148
20.603
29.626
1.352
0.29
1.811
22.179

Input variables: x15Ts (uC), x25NDVI (01), x35local time (decimal), x45Julian day/365 (0
1), x55DEM (km), x65Pw (cm).

In order to analyse the error in more detail, the results for two different land units
were compared, that is forested and non-forested. Better results occur over nonforested land types than over forest land types. SEE and R2 for non-forest were
8.97% and 0.8, compared to values of 11.01% and 0.74 (figure 5) for forest. The
discrepancy is particularly enhanced for measured RH over forest sites, beyond 90%
(figure 5). Statistics for the two land types are listed in table 3. A large difference in
biases for mean, maximum, and minimum values between the two land types was
not achieved. However, better mean deviation (MD) of 2.1% and RMSE of 2.92%
occurred over non-forest than over forest (table 3). These results suggest that using
statistical models adapted to the land types is definitively more appropriate.
However, we found that model accuracy segregated according to the land type (not
shown here) showed no improvement over the single model (equation (8)). The
difference between results over forested and non-forested areas may be explained
from two points of view: canopy top height; topographical variance. The remotely
sensed parameters represent a value at a canopy top height. The canopy top height
of forested (forest canopy height) and non-forested (crop canopy height in this
study) areas is quite different, while the height of RH sensor in the measurement
station is about 1.2 m above ground level. In general, crop canopy is closer to
ground measurement level of RH than forest canopy. In other words, estimation
over crop canopy has a relatively better chance of estimating a parameter at near
surface as compared with estimation over forest canopy. Secondly, topographical

4776

K.-S. Han et al.

Figure 5. Scatterplot of predictions and measurements for development dataset over forest
and non-forest land types (SEE: standard error estimate).

variability in a pixel could also affect the degree of the estimation because this
provides different aspects on the spatial representativeness of a pixel value. The
reasoning associated with topographical variability causing difference in error over
the two land types is discussed in the next section.
5.

Validation and discussion

The goal of the validation procedure is to do an independent test of the accuracy of


the RHe values predicted by equation (8) throughout iterative development and
verification exercises with data subsets. A specific exercise in the validation
procedure consists of evaluating the performance of an estimation equation derived
using part of the data, and applying it to the remainder of the data as an
independent set. Surface conditions, such as vegetation cover and phenology, mainly
vary in seasonal time series. In the validation procedure, the estimation model
established for the development data period (June 1997) was applied to the
validation dataset for July 1997, which is the most adjacent month. This procedure
Table 3. Statistical results for the development dataset for estimating relative humidity (RH)
over forest and non-forest land types.
For development dataset

Forest

Non-forest

Total

Average of measured values (%)


Average of estimated values (%)

51.79
52.20

53.24
52.86

52.54
52.54

Minimum of measured values (%)


Minimum of estimated values (%)

10.95
12.17

16.24
14.71

10.95
12.17

Maximum of measured values (%)


Maximum of estimated values (%)

100.0
93.71

100.0
93.31

100.0
93.71

9.01
11.90

6.91
8.98

7.92
10.48

Mean deviation (%)


RMSE (%)
Number of observations

2060

2223

4283

Estimating near-surface air humidity in eastern Canada

4777

Table 4. Statistical results for the validation dataset for estimating relative humidity (RH)
over forest and non-forest land types.
For validation dataset

Forest

Non-forest

Total

Average of measured values (%)


Average of estimated values (%)

66.76
66.76

67.13
67.13

66.96
66.96

Minimum of measured values (%)


Minimum of estimated values (%)

28.00
26.50

29
28.75

28
26.5

Maximum of measured values (%)


Maximum of estimated values (%)
Mean deviation (%)
RMSE (%)
Number of observations

100.0
107.3
8.16
10.54
1296

100.0
101.5
6.91
8.76
1465

100.0
107.3
7.50
9.63
2761

is also to test the extensibility of the algorithm in the seasonal term. The RH
estimation model established with the development dataset was then evaluated using
the validation dataset for the days between 1 and 10 July 1997.
The validation dataset (2726 cloud-free observations) applied to the estimation
model included the same independent variables and coefficients used in the
development step. RHe values estimated for validation show good agreement with
the measured values. The regression slope was 1.0, the intercept was 0.0000154%,
the R2 was 0.74, and SEE was 9.64%. R2 was slightly lower and SEE decreased as
compared to the development results. Results for non-forest and forest showed the
same tendency as the development dataset. Table 4 and the scatterplots in figure 6
show the better estimation over non-forest lands. Maximum estimation values
reached 100% over both land types and MD and RMSE show almost the same
degree as the development dataset (table 4). The best fit is shown in the estimates for
the validation dataset over non-forest (MD56.91% and RMSE58.76%). In the
summer in Quebec, the RH is high at night and low during the day. The absolute
humidity of the air varies quite slowly on a daily basis, and is a function of the

Figure 6. Scatterplot of predictions and measurements for validation dataset over forest and
non-forest land types (SEE: standard error estimate).

4778

K.-S. Han et al.

broader scale airmass climatology. Relative humidity changes markedly because


of the relationship between temperature and saturation vapour pressure.
Consequently, during the daytime, the vapour pressure deficit increases and the
relative humidity drops. Figure 7 shows a comparison of the RMSE of the
development and validation sets for the diurnal cycle. The measured RH values are
generally higher for the validation period (July) than for the development period
(June). RMSE values for the validation dataset are lower, except for the period
between 14:00 and 21:00 local time. RMSE lines for both datasets show a similar
tendency: RMSE decreases when measured RH is low. The maximum value is 13.9%
at midnight, while the minimum is 6.5% at midday. In short, the degree of error is in
proportion to the degree of measured value. Generally, the validation results show
reasonable agreement with the results for development dataset.
Comparisons between RH and RHe estimates at each station are shown in table 5
(analysed together for both development and validation datasets). R2 values for each
station varied from 0.46 to 0.91 (mean 0.78). The minimum value of RMSE was
7.61% at station 26 (non-forest site); the maximum value was 16.80% at station 2
(forest site). RMSE was generally low when R2 was high. There were 24 stations
with RMSE values lower than the mean (10.16%), and 17 stations with RMSE
values higher than the mean. Figure 8 shows the spatial distribution of the RMSE
values; the spatial distribution of the estimation accuracy is related to land cover
type. RMSE values less than 10.16% (mean RMSE) appeared mainly over nonforest land types (white in figure 8) such as the agricultural regions around Montreal
and St-Jean Lake. An exception was the low RMSE region in the north-west part of
the study area, which was dominated by forest cover. This area has a lower elevation
than other forest areas. On the other hand, RMSE values exceeding 10.16%
occurred in the central, north-east, and south-east parts of the study area, where

Figure 7. Mean RMSE (filled in black) and measured value (filled in white) comparisons
between development and validation datasets.

Estimating near-surface air humidity in eastern Canada

4779

Table 5. Statistical results for the combined dataset (development and validation) of each
station.
Station

Slope

Intercept

R2

RMSE (%)

Number of
observations

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41

0.97
1.40
1.16
0.88
1.09
0.73
0.69
1.37
0.98
1.09
1.17
0.80
1.13
1.01
0.99
0.88
0.87
0.84
0.50
0.91
0.94
1.11
1.04
0.96
1.11
1.06
1.09
0.87
1.15
1.09
1.02
0.98
1.04
0.85
0.82
1.07
0.96
1.00
0.95
0.84
0.85

3.14
218.40
29.26
2.09
23.13
17.49
1.54
224.57
4.58
2.09
21.66
7.36
27.60
20.73
20.24
5.94
8.10
8.26
26.71
11.27
6.49
23.83
22.59
0.89
25.37
25.38
27.53
8.44
25.83
26.83
20.65
3.95
21.37
5.69
6.55
24.36
2.53
4.68
21.86
9.95
5.76

0.81
0.79
0.87
0.72
0.85
0.51
0.46
0.82
0.74
0.84
0.79
0.74
0.91
0.88
0.69
0.67
0.76
0.74
0.41
0.73
0.77
0.78
0.83
0.80
0.81
0.88
0.80
0.76
0.91
0.85
0.79
0.85
0.89
0.75
0.75
0.82
0.80
0.80
0.88
0.80
0.66

8.14
16.80
9.96
12.18
8.64
12.49
13.58
14.09
10.79
9.16
14.11
12.18
8.55
8.37
14.03
9.23
8.07
8.85
15.63
12.44
9.12
10.74
8.37
8.26
10.26
7.61
11.47
8.59
8.47
8.13
10.08
9.36
7.73
10.18
10.37
9.28
8.38
10.00
7.83
8.08
11.14

86
206
211
243
93
96
9
93
66
6
77
190
156
180
114
62
82
88
81
98
94
242
239
239
239
227
218
261
237
233
280
200
254
236
234
236
246
212
220
244
216

Overall

1.00

0.07

0.78

10.16

7044

forest land types dominate (grey in figure 8). In order to enquire into more detail the
spatial distribution of RMSE as shown in figure 8, we analyse with DEM by
dividing into two groups: the group of the stations having lower RMSE than 10.16%
(mean); and the group of the stations having higher RMSE than mean. Table 6
shows the relationship between RMSE and environment by comparing two groups.
Group A includes 28 stations with RMSE values lower than the mean. Seventeen
stations with RMSE values exceeding the mean comprise group B. The majority of

4780

K.-S. Han et al.

Figure 8. Contour map for the spatial distribution of RMSE of measured and estimated
relative humidity. Overlapped image is a simplified land cover map of two classes (forest and
non-forest classes).

group A is non-forest land type; forest land types dominate group B. In addition,
the mean elevation of group B (5248.5 m) is 111.9 m higher than that of group A
(136.6 m). Elevation variability is analysed within a 4 km64 km pixel using a DEM
with 30 m630 m spatial resolution. Elevation variability within a pixel for group B
was four times greater than for group A (table 6). The maximum value for group B is
69.1 m.
The forest-covered zone in the study area is generally distributed over relatively
high mountainous areas where elevation variability within a pixel is high. The
estimation scheme for RH is more suitable for agricultural areas at low elevations
with little intrapixel variability. Nevertheless, RH can be estimated over

Estimating near-surface air humidity in eastern Canada

4781

Table 6. Comparison of environmental characteristics for stations with RMSEs higher and
lower than the all-station average (10.16%).
Group A, (10.16% RMSE
Number of stations
Elevation (m)
maximum
minimum
mean

24

Group B, .10.16% RMSE


17

502.5
13.6
136.6

570.9
37.3
248.5

Elevation variability within a pixel (m)


maximum
18.8
minimum
0.2
mean
4.4

69.1
0.5
16.0

mountainous forest areas. The results indicate that the proposed RH estimation
algorithm could be a useful tool in instances similar to those studied here.
6.

Concluding remarks

The variable RH has been derived here from remotely sensed data (NOAA/AVHRR
and GOES/IMAGER data) and auxiliary spatial and temporal information. Six
input components (Ts, NDVI, Pw, DEM, LT, and JD) are used in a multiple
regression analysis using a stepwise selection function to estimate relative humidity.
The multiple regression polynomial is third-order to match the best fit of the diurnal
variation of Ta and Ts. The relative humidity estimation produced the following
important results: (1) the estimation scheme expressed spatial and temporal
variation on an hourly scale using only one equation; and (2) the values are
directly estimated without the process that is required in the conventional RH
formula to obtain the Td and Ta (equation (5a)), and RMSE is acceptable.
The model for relative humidity had an overall SEE of about 10%. SEE over
forest and non-forest lands was 11.01% and 8.97%, respectively. The results using
this new algorithm are not compared to previous studies, because there have been no
similar investigations; however, a validation dataset yielded similar SEE over forest
(10.54%) and non-forest (8.76%) areas. The results for validation dataset are shown
to be in close agreement with those for development dataset. RMSE differences
between forest and non-forest arise from differences in elevation and spatial
variability. A slightly better fit occurs for non-forest regions that included an
agricultural area. In other words, the estimation is more accurate when a region had
a biologically and geographically homogeneous surface condition within the pixels.
On the other hand, the pixel values over a heterogeneous region can give a mean
value for a corresponding pixel area, but not a representative value. This error could
be reduced if we use only a Sun-synchronous satellite such as NOAA or Terra (EOS
AM-1) on which there is the AVHRR and Moderate Resolution Imaging
Spectroradiometer (MODIS) instruments, respectively, which have much better
spatial resolution than geostationary satellites. This better spatial resolution can
diminish an overload for the spatial representativeness of a pixel value. However, it
should be noted that the temporal resolution of estimation is decreased to the daily
scale because of the polar orbiting system.
Finally, the present algorithm may contribute to higher resolution mapping of
hydro-meteorological elements, such as evapotranspiration and drought or forest

4782

K.-S. Han et al.

fire indices. However, the equations probably cannot be generalized for every
situation, because the derived coefficients are valid only for this special case.
Therefore, future studies should include an analysis of long-term data from various
regions, using the same scheme and input components.
Acknowledgments
This study was supported by funds from SOPFEU. Satellite data were kindly
provided by Environment Canada. Ground-based data were kindly provided by
SOPFEU, Environment Canada, and the POMMIER network. Digital elevation
model data were provided by Geomatic Canada. Finally, the authors would like to
thank anonymous reviewers of this paper. Their comments and suggestions were
appreciated to improve the paper quality.
References
ACHARYA, P.K., BERK, A., BERNSTEIN, L.S., MATTHEW, M.W., ADLER-GOLDEN, S.M.,
ROBERTSON, D.C., ANDERSON, G.P., CHETWYND, J.H., KNEIZYS, F.X., SHETTLE, E.P.,
ABREU, L.W., GALLERY, W.O., SELBY, J.E.A. and CLOUGH, S.A., 1998, MODTRAN
Users Manual, versions 3.7 and 4.0. Air Force Research Laboratory, Hanscom AFB,
MA, USA.
AUMANN, H.H., CHAHINE, M.T., GAUTIER, C., GOLDBERG, M.D., KALNAY, E.,
MCMILLIN, L.M., REVERCOMB, H., ROSENKRANZ, P.W., SMITH, W.L.,
STAELIN, D.H., STROW, L.L. and SUSSKIND, J., 2003, AIRS/AMSU/HSB on the
Aqua mission: design, science objectives, data products, and processing systems.
IEEE Transactions on Geoscience and Remote Sensing, 41, pp. 253264.
CASELLES, V., DELEGIDO, J., HURTADO, E. and SOBRINO, J.A., 1991, Maximum ET
monitoring over La Mancha, Spain, using NOAA-AVHRR data. Proceedings of the
5th International ColloquiumPhysical Measurements and Signatures in Remote
Sensing, Courchevel, France, 1418 January (CD-ROM from ICP).
CASELLES, V., ARTIGAO, M.M., HURTADO, E., COLL, C. and BRASA, A., 1998, Mapping
actual evapotranspiration by combining Landsat TM and NOAA-AVHRR images:
application to the Barrax area, Albacet, Spain. Remote Sensing of Environment, 63,
pp. 110.
CHOUDHURY, B.J., 1996, Comparison of two models relating precipitable water to surface
humidity using globally distributed radiosonde data over land surfaces. International
Journal of Climatology, 16, pp. 663675.
CHOUDHURY, B., DORMAN, T.J. and HSU, A.Y., 1995, Modeled and observed relations
between the AVHRR split window temperature difference and atmospheric
precipitable water over land surfaces. Remote Sensing of Environment, 51, pp.
281290.
CIHLAR, J., LY, H., LI, Z., CHEN, J., POKRANT, H. and HUANG, F., 1997, Multitemporal,
multichannel AVHRR data sets for land biosphere studiesartifacts and corrections.
Remote Sensing of Environment, 60, pp. 3557.
CRESSWELL, M.P., MORSE, A.P., THOMSON, M.C. and CONNOR, S.J., 1999, Estimating surface
air temperatures, from Meteosat land surface temperatures, using an empirical solar
zenith angle model. International Journal of Remote Sensing, 20, pp. 11251132.
CZAJKOWSKI, K.P., MULHERN, T., GOWARD, S.N., CIHLAR, J., DUBAYAH, R.O. and
PRINCE, S.D., 1997, Biospheric environmental monitoring at BOREAS with
AVHRR observations. Journal of Geophysical Research, 102, pp. 29 65129 662.
DRAPER, N.R. and SMITH, H., 1981, Applied Regression Analysis, 2nd edn (New York:
John Wiley).
GEIGER, R., 1965, The Climate near the Ground (Cambridge, MA: Harvard University
Press).

Estimating near-surface air humidity in eastern Canada

4783

GOWARD, S.N. and HOPE, A.S., 1989, Evapotranspiration from combined reflected solar and
emitted terrestrial radiation: preliminary FIFE results from AVHRR data. Advances
in Space Research, 7, pp. 165174.
GOWARD, S.N., CRUICKSHANKS, G.C. and HOPE, A.S., 1985, Observed relation between
thermal emissions and reflected spectral reflectance from a complex vegetated
landscape. Remote Sensing of Environment, 18, pp. 137146.
GUTMAN, G., TARPLEY, D. and OHRING, G., 1987, Cloud screening for determination of land
surface characteristics in a reduced resolution satellite data set. International Journal
of Remote Sensing, 8, pp. 859870.
HAN, K.S., VIAU, A.A. and ANCTIL, F., 2004, An analysis of GOES- and NOAA-derived land
surface temperatures estimated over a boreal forest. International Journal of Remote
Sensing, 25, pp. 47614780.
HESS, S.L., 1959, Introduction to Theoretical Meteorology (New York: Holt, Rinehart &
Winston).
HOPE, A.S. and MCDOWELL, T.P., 1992, The relationship between surface temperature and a
spectral vegetation index of a tallgrass prairie: effects of burning and other landscape
controls. International Journal of Remote Sensing, 13, pp. 28492863.
JONES, P.D., 1995, Land surface temperatures: is the network good enough? Climate Change,
31, pp. 545558.
KARL, T.R., DERR, V.E., EASTERLING, D.R., FOLLAND, C.K., HOFMANN, D.J., LEVITUS, S.,
NICHOLLS, N., PARKER, D.E. and WITHEE, G.W., 1995, Critical issues for long-term
climate monitoring. Climate Change, 31, pp. 185221.
KIDWELL, K.B., 1998, NOAA polar orbiter data users guide service. NOAA/National
Environmental Satellite, Data and Information Service.
MARTIN, M.E., 1997, Mise au point dun indicateur continental devapotranspiration
apparent pour les regions intertropicales Africaines. The`se de matrise, Universite
Laval. Agricultural and Forest Meteorology, 46, pp. 285296.
MONTEITH, J.L. and UNSWORTH, M.H., 1990, Principles of Environmental Physics (New York:
Edward Arnold).
NEMANI, R.R. and RUNNING, S.W., 1989, Estimation of regional surface resistance to
evapotranspiration from NDVI and thermal IR AVHRR data. Journal of Applied
Meteorology, 28, pp. 276284.
NEMANI, R.R., PIERCE, L., RUNNING, S.W. and GOWARD, S.N., 1993, Developing satellitederived estimates of soil moisture status. Journal of Applied Meteorology, 32, pp.
548557.
PAQUET, F., VOGT, J.V. and VIAU, A.A., 1997, Regionalisation of air temperatures using
AVHRR derived surface skin temperatures together with terrain and landcover data.
Proceedings of the International Symposium: Geomatics in the Era of Radarsat
(GER97), Poster, Ottawa, 2530 May, CD-ROM.
PERRY, E.M. and MORAN, M.S., 1994, An evaluation of atmospheric corrections of
radiometric surface temperatures for a semiarid rangeland watershed. Water
Resources Research, 30, pp. 12611269.
PRICE, J.C., 1990, Using spatial context in satellite data to infer regional scale
evapotranspiration. IEEE Transactions on Geoscience and Remote Sensing, 28, pp.
940948.
PRIHODKO, L. and GOWARD, S.N., 1997, Estimation of air temperature from remotely sensed
surface observations. Remote Sensing of Environment, 60, pp. 335346.
PRINCE, S.D. and GOWARD, S.J., 1995, Global primary production: a remote sensing
approach. Journal of Biogeography, 22, pp. 815835.
PRINCE, S.D., GOETZ, S.J., DUBAYAH, R.O., CZAJKOWSKI, K.P. and THAWLEY, M., 1998,
Inference of surface and air temperature, atmospheric precipitable water and vapor
pressure deficit using Advanced Very High-Resolution Radiometer satellite observation: comparison with field observation. Journal of Hydrology, 212/213, pp. 230249.

4784

Estimating near-surface air humidity in eastern Canada

REITAN, C.H., 1963, Surface dew point and water vapor aloft. Journal of Applied
Meteorology, 2, pp. 776779.
RICHTER, R., 1990, A fast atmospheric correction algorithm applied to Landsat TM images.
International Journal of Remote Sensing, 11, pp. 159166.
ROSENBERG, N.J., BLAD, B.L. and VERMA, S.B., 1983, Micro-climate: The Biological
Environment (New York: John Wiley).
SAUNDERS, R.W., 1986, An automated scheme for the removal of cloud contamination from
AVHRR radiances over western Europe. International Journal of Remote Sensing, 7,
pp. 867886.
SAUNDERS, R.W. and KRIEBEL, K.T., 1988, An improved method for detecting clear sky and
cloudy radiances from AVHRR data. International Journal of Remote Sensing, 9, pp.
123150.
SHIN, D., POLLARD, J.K. and MULLER, J.P., 1996, Cloud detection from thermal infrared
images using a segmentation technique. International Journal of Remote Sensing, 17,
pp. 28452856.
SMITH, R.C.G. and CHOUDHURY, B.J., 1991, Analysis of normalized difference and surface
temperature observations over southeastern Australia. International Journal of
Remote Sensing, 12, pp. 20212044.
SMITH, W.L., 1966, Note on the relationship between total precipitable water and surface dew
point. Journal of Applied Meteorology, 5, pp. 726727.
SOBRINO, J.A., COLL, C. and CASELLES, V., 1991, Atmospheric correction for land surface
temperature using NOAA-11 AVHRR channels 4 and 5. Remote Sensing of
Environment, 38, pp. 1934.
TEILLET, P.M., 1992, An algorithm for the radiometric and atmospheric correction of
AVHRR data in the solar reflective channels. Remote Sensing of Environment, 41, pp.
185195.
ULIVIERI, C., CASTRONOUVO, M.M., FRANCIONI, R. and CARDILLO, A., 1994, A split-window
algorithm for estimating land surface temperature from satellites. Advances in Space
Research, 14(3), pp. 5965.
VIAU, A.A., VOGT, J.V. and PAQUET, F., 1996, Regionalisation and mapping of air
temperature fields using NOAA AVHRR imagery. Actes du 9e Congres de
lAssociation quebecoise de teledetection (AQT): La teledetection au sein de la
geomatique, 30 April3 May 1996, Quebec, Canada, CD-ROM.
valuation de la representativite spatiale
VIAU, A.A., VOGT, J.V. and PAQUET, F., 2000, E
thermique des stations meteorologiques du reseau dAndalousie. International Journal
of Remote Sensing, 21, pp. 30833113.
VOGT, J.V., VIAU, A.A. and PAQUET, F., 1997, Mapping regional air temperature fields using
satellite-derived surface skin temperatures. International Journal of Climatology, 17,
pp. 15591579.
WEINREB, M., JAMIESON, M., FULTON, N., CHEN, Y., JOHNSON, J.X., BREMER, J., SMITH, C.
and BAUCOM, J., 1997, Operational calibration of the imagers and sounders on the
GOES-8 and -9 satellites. NOAA/NESDIS/Office of Systems Development.

You might also like