You are on page 1of 24

Journal of Oceanography, Vol. 63, pp.

721 to 744, 2007

Review

Diurnal Sea Surface Temperature Variation and Its Im-


pact on the Atmosphere and Ocean: A Review
Y OSHIMI KAWAI1* and AKIYOSHI WADA2
1
Institute of Observational Research for Global Change, Japan Agency for Marine-Earth Science
and Technology, Natsushima-cho, Yokosuka 237-0061, Japan
2
Meteorological Research Institute, Japan Meteorological Agency,
Nagamine, Tsukuba 305-0052, Japan

(Received 6 November 2006; in revised form 2 April 2007; accepted 14 April 2007)

The importance of the diurnal variability of sea surface temperature (SST) on air-sea Keywords:
interaction is now being increasingly recognized. This review synthesizes knowledge Sea surface
of the diurnal SST variation, mainly paying attention to its impact on the atmosphere temperature,
or the ocean. Diurnal SST warming becomes evident when the surface wind is weak diurnal variation,
intraseasonal
and insolation is strong. Recent observations using satellite data and advanced in-
variation,
struments have revealed that a large diurnal SST rise occurs over wide areas in a
air-sea interaction,
specific season, and in an extreme case the diurnal amplitude of SST exceeds 5 K. The surface flux.
large diurnal SST rise can lead to an increase in net surface heat flux from the ocean
of 5060 Wm2 in the daytime. The temporal mean of the increase exceeds 10 Wm2,
which will be non-negligible for the atmosphere. A few numerical experiments have
indicated that the diurnal SST variation can modify atmospheric properties over the
Pacific warm pool or a coastal sea, but the processes underlying the modification
have not yet been investigated in detail. Furthermore, it has been shown that the
diurnal change of ocean mixing process near the surface must be considered cor-
rectly in order to reproduce SST variations on an intraseasonal scale in a numerical
model. The variation of mixed-layer properties on the daily scale is nonlinearly re-
lated to the intraseasonal variability. The mixed-layer deepening/shoaling process on
the daily scale will also be related to biological and material circulation processes.

1. Introduction (e.g., Reynolds et al., 2002; Rayner et al., 2003; Worley


Sea surface temperature (SST) is the most important et al., 2005). Although these high-quality datasets enable
factor in air-sea interaction. The sea surface is the lower us to investigate long-term and large-scale variations, the
boundary of the atmosphere, and SST influences weather spatial and temporal resolution of these SST data is too
and climate. On the other hand, SST is also controlled by poor to resolve eddies and temporal high-frequency vari-
atmospheric conditions. Accurate SST data are indispen- ation. Several researchers have shown that the imposi-
sable for climate monitoring, prediction and research. SST tion of an hourly surface forcing is essential to reproduce
is also important in chemical and biological oceanogra- diurnal and intraseasonal SST variations in a numerical
phy. In order to clarify the mechanism of the global cli- model (e.g., Weller and Anderson, 1996; Sui et al., 1997b;
mate system we need a high-quality SST dataset and Shinoda and Hendon, 1998; Bernie et al., 2005). Further-
knowledge of air-sea interaction processes. more, Li et al. (2001) and Clayson and Chen (2002) indi-
SST data have been collected for more than a cen- cated that the diurnal SST variation can have an impact
tury, and form the most abundant dataset in oceanogra- on the atmosphere over the western Pacific warm pool.
phy. Nowadays several in situ or analytical long-term glo- The air-sea interaction on the daily scale may play an
bal SST datasets are produced and released to the public important role in the Madden-Julian Oscillation (MJO),
and in turn the El Nio and Southern Oscillation (ENSO)
* Corresponding author. E-mail: ykawai@jamstec.go.jp and climate (cf. Slingo et al., 2003; Dai and Trenberth,
CopyrightThe Oceanographic Society of Japan/TERRAPUB/Springer 2004).

721
Diurnal variation, which is caused by solar radiation called the warm layer (Fairall et al., 1996). The
and the earths rotation, is one of the dominant variations thermocline near the surface that develops only in the
in SST. The existence of the diurnal variation in SST was daytime is referred to as diurnal thermocline. Further-
known at least one century ago (cf. Sverdrup et al., 1942; more, at the top of the ocean, a very thin cool layer, which
Roll, 1965). Sverdrup et al. (1942) indicated that the is usually called the thermal skin layer, cool skin layer
diurnal variation of sea temperature in general is so small or simply skin layer, almost always exists (e.g.,
that it is of little importance to physical and biological Saunders, 1967; Katsaros, 1980; Robinson et al., 1984;
processes in the sea, but knowledge of the small varia- Ward and Donelan, 2006). The thickness of this layer is
tions is essential to the study of the diurnal exchange of usually the order of 0.11 mm, and the temperature at the
heat between the atmosphere and the sea. The diurnal top of the skin layer is generally several tenths of a de-
amplitude of SST is O (0.1 K) on average, but often gree colder than the temperature below it because eddy
reaches a few degrees and can exceed 5 K in extreme diffusion becomes less than molecular diffusion just close
cases (e.g., Flament et al., 1994; Yokoyama et al., 1995). to the surface. The phenomenon of the temperature drop
The satellite remote sensing community has clearly rec- occurring in this thin layer is called the skin effect.
ognized that the diurnal SST variability has to be ad- While the diurnal thermocline vanishes by sunrise next
equately considered to provide better accuracy of satel- morning, the skin layer usually exists in both the daytime
lite-derived SST (e.g., Hepplewhite, 1989; Wick et al., and nighttime, even in windy conditions (Donlon et al.,
2002; Donlon and the GHRSST-PP Science Team, 2005; 2002). In the daytime, the temperature difference across
Notarstefano et al., 2006). However, the diurnal varia- the skin layer becomes smaller due to the absorption of
tion has often been neglected in analytical SST datasets shortwave radiation in this layer. The temperature at the
or numerical modeling, and the effects of air-sea interac- top can theoretically even become higher than that at the
tion on the daily scale have not yet been widely recog- bottom when the air temperature is very much higher than
nized nor adequately revealed. that of the water, as has been confirmed by a tank experi-
Many researchers have recently become interested ment (Ward and Donelan, 2006). The skin layer thick-
in the diurnal SST variation, and increasing numbers of ness and the diurnal thermocline depth vary with surface
studies pay attention to air-sea interaction on the daily heat and momentum fluxes (e.g., Fairall et al., 1996; Ward
scale. This review synthesizes knowledge of the diurnal and Donelan, 2006). A better knowledge of SST demands
SST variation, focusing on why and how the diurnal vari- consideration of both the diurnal warming near the sur-
ation is important for the atmosphere and the ocean. face and the effect of the skin layer.
Soloviev and Lukas (2006) gave a detailed, comprehen- A sharp temperature gradient sometimes appears
sive account of the structure and dynamics of the near- above 1-m depth in the daytime (e.g., Yokoyama et al.,
surface ocean. Furthermore, previous studies of the skin 1995; Soloviev and Lukas, 1997; Ward, 2006; see Figs.
effect (see Section 2) were reviewed or introduced by 13 in the present paper). The large temperature differ-
Katsaros (1980), Robinson et al. (1984), and Ward et al. ence between the sea surface and about 1-m depth, where
(2004a), for example. In this paper we focus on the diur- ships and buoys usually measure the seawater tempera-
nal sea surface warming and its effect on the atmosphere, ture as SST, has been recognized as one of the major
with less emphasis on the skin effect. We do not take up sources of error in satellite-derived SST. The satellite in-
the diurnal variation caused by the tide here. frared sensor, microwave sensor, and in situ sensors ob-
First, the vertical temperature structure near the sur- serve different sea surfaces, i.e., skin, subskin, and a
face and the definition of SST are explained in Section 2. depth of one or a few meters, respectively. Hence it was
Section 3 summarizes previous observational studies that indispensable to define SST exactly and consider careful
showed large diurnal variations in SST. Section 4 intro- treatment of different SSTs when producing an accurate
duces several models used to simulate the diurnal SST SST dataset. The Global Ocean Data Assimilation Ex-
variation. Sections 5 and 6 discuss how the diurnal vari- periment (GODAE) High-Resolution SST Pilot Project
ability of SST affects the surface flux estimation and the (GHRSST-PP) Science Team has been doing work on
atmosphere. Some main issues on modeling and observa- coordinating a new generation of global, multi-sensor,
tion of the diurnal SST variation are mentioned in Sec- high-resolution SST products for the benefit of the op-
tion 7. The paper is summarized in Section 8. erational and scientific community (Donlon et al., 2007).
This Science Team has defined five kinds of SST:
2. Vertical Thermal Structure near the Sea Surface interface SST (SSTint), skin SST (SSTskin), subskin SST
and SST Terminology (SSTsubskin), sea temperature at depth (SSTdepth), and foun-
Under clear and calm conditions, thermal stratifica- dation SST (SSTfnd) (Donlon and the GHRSST-PP Sci-
tion is formed within the top few meters of the ocean due ence Team, 2005). A schematic picture of the vertical tem-
to strong insolation. This diurnal stratified layer is often perature profile is shown in Fig. 1, and the definitions of

722 Y. Kawai and A. Wada


Table 1. Definitions of the five kinds of SST proposed by Donlon and the GHRSST-PP Science Team (2005) and Donlon et al.
(2007). See also Fig. 1.

Name Abbreviation Temperature represented Instrument to measure it


Interface SST SSTint Theoretical temperature at the precise air-sea None
interface
Skin SST SSTskin Temperature within the conductive diffusion- Infrared radiometer operating in the 3.712
dominated sub-layer at a depth of micrometer spectral waveband
approximately 1020 m
Subskin SST SSTsubskin Temperature at the base of the conductive laminar Microwave radiometer operating in the 611 GHz
sub-layer frequency range, high-performance autonomous
profiler (SkinDeEP, Ward et al., 2004b)
Sea temperature at depth SSTdepth In situ temperature measured below the Traditional in situ sensor (thermistor, CTD, XBT,
conductive laminar sub-layer, which is etc.)

Foundation SST SSTfnd Temperature of the water column free of diurnal


temperature variability or equal to the SSTsubskin
in the absence of any diurnal signal

the SSTs are explained in Table 1. In actuality, we cannot SST (Notarstefano et al., 2006). We should note that the
know SSTint even with current technology (Donlon et al., satellite SST tuned against SSTdepth has the above char-
2002), and SSTskin is usually utilized as a substitute for acteristics. For the recent satellite SST products from the
SSTint on the assumption that SSTskin is close enough to Along Track Scanning Radiometer (ATSR) (Mutlow et
the true SSTint. The in situ SST measured at about 1-m al., 1994) and the Tropical Rainfall Measurement Mis-
depth or deeper has been called bulk SST. The Science sion (TRMM) satellites Microwave Imager (TMI)
Team recommends using SSTdepth rather than the con- (Gentemann et al., 2004), the algorithms were developed
ventional term bulk SST referring to an in situ SST based on the radiance simulated by a radiative transfer
measurement made at 1-m depth as SST1m, for example. model in order to derive SSTskin or SST subskin exactly.
This terminology is introduced to encourage reporting of The atmosphere senses only the exact interface be-
the measurement depth along with the temperature, be- tween the atmosphere and the ocean. Hence we have to
cause, as depicted in Fig. 1, the temperature can change know SSTint (or SSTskin practically) and its diurnal varia-
drastically with depth when the diurnal thermocline is tion for accurate estimation of air-sea heat and gas fluxes
formed. The new concept of foundation SST is intro- (e.g., Sarmiento and Sundquist, 1992; Fairall et al., 1996).
duced as a more precise, well-defined quantity instead of If the temperature at a few meters depth is used as SSTint
previous, loosely-defined bulk SST, which is affected in a flux calculation, the atmosphere will receive incor-
by the diurnal warming. SST fnd will be similar to a rect heat and water vapor from the ocean. This impact is
nighttime minimum or pre-dawn value at depths of ~15 not always negligible, as discussed in Sections 5 and 6.
m, but note that this depth is only a rough estimate and
can deviate from this range in some cases. This paper 3. Observational Facts Concerning Diurnal SST
basically adopts the terminology proposed by Donlon and VariationWhen and How can the Diurnal Vari-
the GHRSST-PP Science Team (2005). When referring ation Become Large?
to the temperature near the surface in a general sense,
loosely, the authors simply use SST. 3.1 In situ observation
While a satellite sensor sees SSTskin or SST subskin, in Sverdrup et al. (1942) and Roll (1965) introduced
situ SST observed from ships and buoys is the tempera- some early observational studies of diurnal SST varia-
ture at about 1-m depth or deeper. Algorithms of satel- tion. These early studies used research vessel SST data,
lite-derived SST are conventionally tuned by using buoy- which correspond to SSTdepth, and the depths of the meas-
observed SST, i.e., SSTdepth , as the truth. Hence the aver- urements were not mentioned. For example, according to
age of the satellite SST agrees with that of SSTdepth. How- them, a 1923 report based on data of German and British
ever, it is expected that the variability of the satellite SST research vessels obtained during 18721903 stated that
reflects that of SSTskin or SST subskin, rather than SST depth mean diurnal amplitude of SST in the low latitude was
(cf. Kearns et al., 2000; Kilpatrick et al., 2001; Wick et about 0.30.4 K, and the amplitude decreased to 0.26 K
al., 2002; Stuart-Menteth et al., 2003; Dong et al., 2006). at 4555S and 0.15 K at 5560S. Koizumi (1956)
Some researchers call such satellite SST pseudo-bulk analyzed ocean station data at Extra (39N, 153E) and

Diurnal SST Variation and Its Impact 723


Tango (29N, 135E) in the northwestern Pacific. He ligible influence on air-sea heat flux estimation (see Sub-
showed that the diurnal amplitude of SST was smallest section 5.1). According to the results of Soloviev and
(0.15 K at Extra and 0.36 K at Tango) in winter and larg- Lukas (1997) and Donlon et al. (2002), the diurnal
est in summer (0.52 K at Extra and 0.65 K at Tango) on a thermocline almost disappears when wind speed exceeds
monthly average. Koizumi (1956) also indicated that in about 5 m s1. Furthermore, Webster et al. (1996) and
summer the daily maximum occurred later (1500 LST) Soloviev and Lukas (1997) indicated that the shallow
than in winter (around 1300 LST), although the halocline caused by precipitation also affects the diurnal
seasonality of the time of the daily minimum was not SST variation.
obvious. Stommel and Woodcock (1951) and Stommel et Infrared radiometers are sometimes operated on re-
al. (1969) reported examples of large diurnal SST rise search vessels or an oil derrick to measure SSTskin (e.g.,
reaching 11.5 K in the Gulf of Mexico and south of Ber- Schlessel et al., 1987, 1990; Fairall et al., 1996; Donlon
muda in spring. et al., 1998; Kearns et al., 2000; Barton et al., 2004; Nicls
As mentioned above, in the early and mid twentieth et al., 2004). SSTskin is now urgently required for satel-
century it was indicated that the diurnal amplitude of SST lite SST tuning/validation and the study of air-sea fluxes
was about 0.20.6 K on average, varying with latitude and interaction, rather than SSTdepth. SSTskin shows larger
and season. It was also known that the diurnal amplitude diurnal variations than SSTdepth under calm and clear con-
depended on cloudiness and wind speed, and could reach ditions (e.g., Fairall et al., 1996; Donlon et al., 1998;
about 1.5 K on clear, calm days. Stronger winds induce Clayson and Chen, 2002). During windy conditions, the
stronger turbulent mixing in the ocean and prevent ther- variation of SSTskin is almost the same as that of SSTdepth,
mal stratification. Furthermore, stronger turbulence in the and SSTskin is a little cooler than SSTdepth and SSTsubskin
atmosphere draws heat from the ocean. These two func- due to the skin effect and the absence of the diurnal
tions mean that the diurnal amplitude of SST decreases thermocline.
as the wind becomes stronger. On the other hand, stronger Several kinds of infrared radiometer for in situ
insolation causes a greater diurnal SST rise due to the SSTskin observations have been developed, such as M-
absorption of radiation near the sea surface. About 60% AERI, ISAR, SISTeR, CIRIMS, etc. (cf. Barton et al.,
of incoming shortwave radiation is absorbed within the 2004). However, infrared radiometers that can measure
upper 1 m of the ocean (cf. Soloviev and Lukas, 2006). SSTskin with high accuracy are generally so expensive and
In relatively recent years, larger diurnal SST varia- complicated that the radiometric SST observations are
tions have been reported, using advanced instruments. now restricting their usage to a limited number of research
Bruce and Firing (1974) showed an example of a diurnal vessel cruises. Wide spread usage of tough, low-cost ra-
temperature rise exceeding 3 K in the layer of 0~1-m diometers will be also necessary (cf. Donlon et al., 1998).
depth, using slow-sinking Expendable Bathythermograph
(XBT) probes. Other researchers also observed diurnal 3.2 Satellite observation
amplitudes of SST depth or SST skin reaching 23 K or more Operational satellite SST observations started in the
under calm and clear conditions with buoys, profiling 1980s, and have provided the ability to investigate diur-
floats, and infrared radiometers on vessels (e.g., Price et nal SST variation over a wide region. Deschamps and
al., 1986, 1987; Yokoyama et al., 1995; Weller and Frouin (1984) studied the diurnal heating of the sea sur-
Anderson, 1996; Webster et al., 1996; Soloviev and Lukas, face in the Mediterranean Sea using the SST observed by
1997; Kawai and Kawamura, 2002; Ward, 2006). In par- the Heat Capacity Mapping Radiometer (HCMR) satel-
ticular, diurnal SST variations were minutely observed lite. They indicated that the day-night SST difference
in the Pacific warm pool during the Tropical Ocean Glo- clearly depended on wind speed, and exceeded 3 K in
bal Atmosphere (TOGA)/Coupled Ocean-Atmosphere very calm cases. Cornillon and Stramma (1985) and
Response Experiment (COARE) (cf. Godfrey et al., 1998). Stramma et al. (1986) used the SST derived from the
For example, Soloviev and Lukas (1997) observed many National Oceanic and Atmospheric Administration
diurnal thermocline profiles using a special instrument (NOAA) satellite/Advanced Very High Resolution Radi-
called a free-rising profiler during COARE (Fig. 2). The ometer (AVHRR) observation to show that the SST often
depth of the diurnal thermocline tends to be shallower as became much higher in the daytime than the nighttime in
the gradient of the thermocline increases, and the forma- the northern Atlantic. The large diurnal warming occurred
tion of a very sharp diurnal thermocline is restricted within around the ridge of the Azores-Bermuda high pressure in
0~1-m depth. Ward (2006) also showed temperature strati- spring and summer, and the maximum day-night differ-
fication of up to 2.7 K formed above 1-m depth using an ence reached 34 K. Flament et al. (1994) reported an
autonomous profiling float called the Skin Depth Ex- example of diurnal SST amplitude reaching 6.6 K locally
perimental Profiler (SkinDeEP) (Fig. 3). This large tem- off California using the AVHRR SST. Diurnal SST varia-
perature difference across the warm layer has a non-neg- tion is also affected by biological process, and it can be

724 Y. Kawai and A. Wada


Fig. 1. Schematic showing (a) idealized nighttime vertical tem-
perature deviations from the foundation SST and (b) ideal-
ized daytime vertical temperature deviations from the foun-
dation SST in the upper ocean. From Donlon and the
GHRSST-PP Science Team (2005). Courtesy of C. J.
Donlon.

Fig. 3. Temperature-depth measurements from SkinDeEP at


22.52N, 109.59W on 10 Oct. 1999 (graph I). Wind speed
(u) and downwelling shortwave radiation (Q sw) (graph II).
Temperature differences: SST skin-SST subskin (blue) and
SST subskin-SSTdepth (red) (graph III). Heat loss differences:
Q(SST skin )-Q(SST subskin ) (blue) and Q(SST subskin )-
Q(SST depth) (red) (graph IV). Q(SSTskin ), Q(SSTsubskin ), and
Q(SST depth) are the surface net heat flux calculated by us-
ing SST skin, SST subskin, and SST depth as SSTint, respectively.
SST skin was measured with an infrared radiometer. From
Ward (2006), Copyright 2006 American Geophysical Un-
Fig. 2. Vertical temperature profiles in the TOGA COARE do- ion. Reproduced by permission of American Geophysical
main obtained by a free-rising profiler during different wind Union.
speed conditions taken at approximately the same afternoon
time on different days. Reprinted from Soloviev and Lukas
(1997), Copyright 1997, with permission from Elsevier.
morning and afternoon in the Japan Sea. The regional
satellite SST dataset produced by Sakaida et al. (2000)
was used here. In this case the diurnal SST rise exceeded
detected by satellite observation. Kahru et al. (1993) re- 3 K off the coast of northern Japan, and clearly corre-
ported that the AVHRR SST increased locally by up to sponded to the weak wind. In the Japan Sea the diurnal
1.5 K in the daytime in the southern Baltic Sea, corre- warming often becomes quite large under a ridge of high
sponding to surface accumulations of cyanobacteria. pressure or behind the high mountains of Japan.
Large diurnal SST variations also frequently occur Stuart-Menteth et al. (2003) first showed the global
in the marginal seas around Japan from spring to summer climatological distribution of the day-night SST differ-
(Yokoyama et al., 1995; Kawai and Kawamura, 1997, ence using six years of AVHRR SST data (see their fig-
2000, 2002; Kawai et al., 2006a). Figure 4 shows an ex- ure 9). From spring to summer the difference becomes
ample of the extremely large SST difference between larger in the Mediterranean Sea, the Bay of Bengal, the

Diurnal SST Variation and Its Impact 725


Fig. 4. NOAA/AVHRR 0.01-grid SST images at (a) 0645 LST (2145 UT) and (b) 1501 LST (0601 UT) on 27 July 1999. (c) SST
difference between (a) and (b). (d) Daily mean 0.25-grid surface wind speed observed with QuikSCAT/SeaWinds on the
same day.

Fig. 5. Seasonal mean day-night difference of the AMSR-E ver. 5 SST produced by Remote Sensing Systems during June 2002
May 2006. Original grid size of the SST is 0.25, and the seasonal mean is calculated in 1 grids. Nominal observation time is
approximately 0130/1330 LST.

726 Y. Kawai and A. Wada


Arabian Sea, the seas around Japan, the north Pacific off from boreal winter to spring. This is consistent with the
North America, and the Azores-Bermuda high pressure analysis of buoy data by Cronin and Kessler (2002). Ac-
belt. Basically, the diurnal warming is large in the tropics cording to Clayson and Weitlich (2005, 2007), this pat-
through the year. The diurnal SST variation was also stud- tern disappeared in the mature phase of the 199798
ied using geostationary satellite data (Wu et al., 1999; ENSO event. An ENSO event changes the spatial and tem-
Tanahashi et al., 2003). These authors also reported cases poral variations of the diurnal SST amplitude as a result
when the day-night SST difference reached a few degrees of the characteristically different surface conditions as-
in the Azores-Bermuda high or the seas around Japan. sociated with ENSO in this region (Cronin and Kessler,
Gentemann et al. (2003) showed that the TMI SST 2002; Kawai and Kawamura, 2005). Using satellite data,
data could detect the diurnal variation, which clearly de- Kawai and Kawamura (2005) and Clayson and Weitlich
pended on solar radiation and wind speed. The micro- (2007) also indicated that the diurnal SST amplitude in
wave sensor has a great advantage in studying SST vari- the western equatorial Pacific varies in association with
ations because it can observe both SST and wind speed MJO.
synchronously through clouds. On the other hand, the
infrared sensor on the geostationary satellite has another 4. Modeling Diurnal Variations near the Surface
advantage in that it measures SST hourly with a higher Many kinds of numerical or empirical models have
spatial resolution, although infrared radiation cannot pen- been used to investigate the diurnal variations of the up-
etrate cloud. per ocean. However, no model can simulate the diurnal
The Advanced Microwave Scanning Radiometer for variations perfectly (Soloviev and Lukas, 2006). In gen-
the Earth Observing System (AMSR-E) on the Aqua sat- eral, numerical models concerned with simulations of the
ellite has been operating since June 2002 (e.g., Dong et diurnal variations are roughly categorized as diffusion-
al., 2006). AMSR-E is the first microwave sensor that type, bulk- or slab-type, and transilient models. Soloviev
can observe SST all over the oceans. We calculated sea- and Lukas (2006) also provided general information about
sonal mean day-night differences from the AMSR-E ver- the modeling of the diurnal variations. In addition, they
sion-5 SST dataset produced by Remote Sensing Systems. showed a simulated diurnal variation in the Pacific warm
The characteristics of the spatial distribution and seasonal pool by a transilient model. This section reviews numeri-
variation of the day-night difference shown in Fig. 5 are cal models for the simulation of the diurnal variations,
basically consistent with the results of Stuart-Menteth et except the transilient model described by Soloviev and
al. (2003). The spatial and seasonal variations of the day- Lukas (2006).
night difference shown by Stuart-Menteth et al. are not
as clear as those in our Fig. 5, especially in the southern 4.1 Diffusion models
hemisphere, mainly due to the lack of the AVHRR SST Diffusion models are subdivided into the following
sampling hindered by clouds. categories: (1) which parameterize the turbulent mixing
In boreal spring before the Indian monsoon, the day- and eddy-diffusion directly in empirical or semi-empiri-
night SST difference is greater than 0.5 K over the north- cal ways; (2) which estimate turbulence quantities by tur-
ern Indian Ocean. A large day-night difference in the high bulent closure at each level. The former, based on our
latitude of the northern hemisphere, and in the subtropi- knowledge of the Monin-Obukhov similarity theory in
cal high pressure belt of the southern hemisphere in sum- the surface boundary layer, are the models proposed by
mer is also clearly captured. Around 45S in the Indian Kondo et al. (1979) and Large et al. (1994). The latter
Ocean and the Pacific Ocean sections the mean day-night are those of Mellor and Yamada (MY) (1982), and modi-
difference is less than 0.2 K throughout the year due to fied version of MY (e.g., Kantha and Clayson, 1994).
the strong westerlies. Interestingly, in the high-latitude Kondo et al. (1979) proposed a simple one-dimen-
region around 60S the mean day-night difference reaches sional model to reproduce observed behavior of the sur-
0.30.5 K in austral summer. The diurnal SST variability face current and daily sea temperature within 10-m depth
around the Antarctic has rarely been reported. This needs near the sea surface. Their model specializes in simulat-
to be investigated in future study. ing the surface boundary layer. Results of the numerical
In the narrow zonal area in the eastern equatorial experiment suggested that the diurnal range of upper-layer
Pacific west of the Galapagos Islands, the day-night dif- temperature was certainly dependent on wind speed. The
ference becomes notably large, reaching a maximum in diurnal amplitude of simulated SST reached its peak of
boreal spring and a minimum in boreal autumn. Deser nearly 2 K when the 10 m-height wind speed was 2.5
and Smith (1998) and Clayson and Weitlich (2005, 2007) ms1 on a clear equinox day at 35N. They also indicated
also indicated that the diurnal SST amplitude has a local that the more stable the sea surface layer was, the faster
maximum over the cold tongue in the eastern equatorial the surface drift current, due to inhibition of downward
Pacific. The diurnal amplitude in this area becomes larger momentum transfer.

Diurnal SST Variation and Its Impact 727


Kawai and Kawamura (2000) modified Kondo et al.s effects strengthen the vertical mixing. Recently, efforts
model by replacing dimensionless shear functions to make have been made to incorporate these effects into the mixed
turbulent transfer larger for stable cases. Kawai and layer model (e.g., Noh et al., 2004), which will improve
Kawamura (2000) were able successfully to simulate a the reproduction of the near-surface temperature varia-
very sharp diurnal thermocline within 1-m depth observed tion.
in Mutsu Bay (Yokoyama et al., 1995) by using both the b. Simplified model
modified version of Kondo et al.s model and the sec- If we desire to simulate sharp diurnal thermocline
ond-order turbulence-closure model developed by MY and diurnal SST variations in numerical models, we need
(Fig. 6). a fine vertical grid interval near the surface (Bernie et
al., 2005). This requires an enormous computational load.
4.2 Bulk or slab models In order to avoid it, some simplified models have been
Modeling under the assumption of a constant profile proposed that specialize in simulating the diurnal varia-
of sea temperature, salinity and current within a mixed tion near the surface. Fairall et al. (1996) proposed a
layer is in general categorized as a bulk or slab model greatly simplified form of the PWP model that ignores
(hereafter, bulk model). The bulk model is subdivided into full mixed-layer dynamics in order simply to predict the
integral and layer types. Integral models represent turbu- diurnal variation of the near-surface temperature only.
lent mixing at the base of the mixed layer by entrainment. They assumed that the temporal integrals of surface heat
Some of the entrainment rate parameterizations are de- and momentum fluxes are confined within the warm layer,
rived from the kinetic energy equation under the condi- and the temperature profile in the warm layer is linear
tion that the energy equation is conserved. In general, a (Fig. 7(a)). The depth of the warm layer is determined by
velocity jump has to be assumed at the bottom of the requiring that the bulk Richardson number is no greater
mixed layer in the integral models. Layer models use a than a critical number. Their warm layer model is well-
modeling methodology that estimates turbulent kinetic known and is often utilized for air-sea flux estimation in
energy in every vertical layer. Simplified versions of the combination with their cool-skin model.
latter models have been developed recently. We here di- The Fairall et al. model is simple and convenient,
vide bulk models into multilayer type and simplified sin- but the linear temperature profile does not always agree
gle layer type. We do not refer to the integral type in the with observations (e.g., Ward et al., 2004a). Zeng and
present review. Beljaars (2005) developed another simple scheme to es-
a. Multilayer model timate SSTskin by assuming a more realistic profile shape
A bulk model proposed by Price et al. (1986) has (Fig. 7(b)). The temperature profile in the warm layer T(z)
been frequently applied to studies of the diurnal SST vari- is given by the following formula:
ation. The model is often called the Price-Weller-Pinkel
(PWP) model. Its numerical scheme is based on a dy-
z
namic instability model (DIM) (Price et al., 1978). The
concept of DIM is that the rate of change of potential
T ( z ) = SSTsubskin
DT
[ ]
SSTsubskin T ( DT ) , (1)
energy during deepening of the mixed layer balances that
rate of energy released from the mean flow by the reduc- where z is the depth, is the depth of the skin layer, and
tion of vertical shear. The PWP model is a modified ver- DT is the depth of the warm layer. An empirical param-
sion of the DIM by Price et al. (1978), including a mix- eter is assumed to be 1 in the Fairall et al. model and
ing process in the stratified part below the mixed layer. 0.3 in the Zeng-Beljaars model. Zeng and Beljaars (2005)
It is worth noting, though, that the diurnal SST am- mentioned that the diurnal variation of ocean tempera-
plitude simulated by the PWP model tends to be too large ture usually becomes small enough at 2~4-m depth, which
when compared with observations (Kantha and Clayson, corresponds to SSTfnd mentioned in Section 2. They de-
1994; Large et al., 1994). Large et al. (1994) indicated termined D T in advance empirically within this depth
that this disagreement of the PWP model could be caused range. The eddy diffusion coefficient in the surface layer
by insufficient vertical mixing. SST is more sensitive to was determined on the basis of the Monin-Obukhov simi-
the strength of the vertical thermal diffusion when the larity theory.
surface stratified layer becomes less than 1-m thick un- For the purpose of reproducing the diurnal cycle of
der calm and clear conditions. The representation of the SST in ocean general circulation models (OGCM),
vertical mixing in the stable stratified layer is one of the Schiller and Godfrey (2005) proposed a simple method
crucial keys to predict diurnal SST variation accurately, that incorporates an extra variable-depth diurnal sublayer
but is still imperfect. The mixed layer models mentioned in the top model layer (Fig. 7(c)). The sublayer exists
here do not take the turbulence generated by wave break- only when the total buoyancy received in the top layer is
ing and Langmuir circulation into consideration. These positive. Their sublayer model is based on the concept of

728 Y. Kawai and A. Wada


where PS is the daily peak surface solar radiation
(W m2), P is the daily mean precipitation rate (mm h1),
and U is the daily mean wind speed (m s 1). a, b, c, d, e,
and f are the coefficients (Table 2). The estimated SSTskin
is shown in Fig. 8. If there is precipitation, SSTskin be-
comes higher due to the saline stratification. The SSTskin
estimated by Webster et al.s empirical model cannot ex-
ceed 3 K, even in the extremely calm and strong insola-
tion case. This upper limit seems too small in compari-
son with the observational studies mentioned previously.
They produced the model (2) based on the simulation re-
sults of Kantha and Claysons (1994) mixed layer model
under various forcing conditions. The reason for the
smaller upper limit of model (2) may be that the sensitiv-
ity of the mixed layer model to the forcing was imper-
fect.
Fig. 6. Vertical temperature profiles in Mutsu Bay at 1430 LST The effects of precipitation and latent heat flux will
on 7 July 1992. Solid line and open circles represent ob-
certainly be needed for better estimation of the diurnal
served values. Broken line and asterisks represent tempera-
ture simulated by the second-order turbulence-closure model
amplitude of SST (e.g., Price et al., 1987; Soloviev and
of Mellor and Yamada (1982). Chain line and pluses repre- Lukas, 1997; Kawai and Kawamura, 2003). However, they
sent those simulated by Kawai and Kawamuras (2000) have a secondary effect on the diurnal SST variation, and
model. Reproduced from Kawai and Kawamura (2000). the diurnal amplitude can even be approximately evalu-
ated from wind speed and solar radiation only (Kawai
and Kawamura, 2002; Gentemann et al., 2003; Clayson
and Weitlich, 2005). In reality it is not easy to know daily
precipitation rate accurately over a wide ocean region.
DIM, like that of Fairall et al. (1996). Schiller and Stuart-Menteth et al. (2003) estimated the diurnal ampli-
Godfreys model could well reproduce the observed di- tude of SST over the globe using Kawai and Kawamuras
urnal cycle of 2.5-m-depth temperature. However, note (2002) empirical formula, which uses only wind speed
that their model does not calculate SSTskin, and the tem- and solar radiation, and confirmed that the spatial distri-
perature in the sublayer is assumed to be independent of bution of the estimated diurnal amplitude agreed well with
depth. Hence the SST simulated by their model is expected that of the day-night satellite SST difference.
to be a little different from SSTskin. Furthermore, we need While the above models can estimate only the am-
to bear in mind that the simplification of the models in- plitude, which is defined as the daily maximum-minimum
troduced in this subsection would neglect some dynami- difference, Zeng et al. (1999) and Gentemann et al. (2003)
cal processes in the upper ocean, which may affect air- proposed simple empirical models that can estimate the
sea interaction. diurnal SST rise from the morning minimum at each hour.
Clayson and Curry (1996) also developed a method to
4.3 Empirical parametric models estimate the diurnal SSTskin cycle. They assumed a co-
The diurnal SST variation depends primarily on wind sine-shaped diurnal cycle with the amplitude estimated
speed and solar radiation, so the diurnal SST amplitude by Webster et al.s empirical model (2), and added it to
can be estimated from these meteorological data. Price et daily predawn SSTskin, which was determined from satel-
al. (1987) proposed an empirical model to evaluate the lite-derived SSTskin data by linear interpolation. Further-
diurnal amplitude of 0.6-m-depth temperature. They re- more, Li et al. (2001) proposed a simple technique to es-
lated the amplitude of SST0.6m with surface stress and timate SSTskin at each hour in a numerical model by ap-
surface heat flux. Another simple model proposed by plying Webster et al.s model. Li et al. (2001) assumed
Webster et al. (1996) included precipitation rate as well that the variation of SSTskin was determined by the sur-
as wind speed and solar radiation based on mixed-layer face wind and solar radiation one hour previously, and
model simulations during COARE. Their empirical model the hourly anomaly of SSTskin from its daily mean at the
of the diurnal amplitude of SSTskin (SSTskin) has the fol- i-th hour on the j-th day (T i,j) could be expressed in the
lowing form: following form:

SSTskin
1 24 12 i + 1
= f + a( PS) + b( P) + c[ln(U )] + d ( PS) ln(U ) + e(U ), Ti, j = i 1, j i , j 1 1 , (3)
(2 ) i =1
12 12

Diurnal SST Variation and Its Impact 729


Fig. 7. Schematics of vertical temperature profile near the surface assumed in (a) Fairall et al.s (1996) model, (b) Zeng and
Beljaars (2005) model, and (c) Schiller and Godfreys (2005) sublayer scheme. DT is the depth of the warm layer (sublayer).
Z k=1 in (c) is the thickness of the top layer of an ocean mixed model. is the depth of the skin layer. T top and T bot(z) in (c) are
the temperatures in the sublayer, and the layer under the sublayer, respectively. Tbot(z) may depend on depth, but is not solved
in the scheme. In (a) and (c) DT is time-dependent, while it is given in advance empirically in (b).

( ) ( )
i, j = f + a Si, j + c ln Ui, j Table 2. Coefficients for the empirical formula (2). Reproduced
from Webster et al. (1996).
Si. j > 0
+ d ( Si, j ) ln(Ui, j ) + e(Ui, j ), ( 4)
i, j = 0, Si. j 0, Coefficient U > 2 m s1 U 2 m s1
f 0.262 0.328
a 0.00265 0.002
where Si,j and Ui,j is the hourly surface solar radiation b 0.028 0.041
and wind speed at the i-th hour on the j-th day, respec- c 0.838 0.212
tively. The precipitation is ignored here. The time of lo- d 0.00105 0.000185
24 e 0.158 0.329
cal sunset is defined as i = 1, and i , j 1 is the sum of
i =1
on the previous day. They showed that their
parameterization could well reproduce diurnal SST skin
variations. Since these simple schemes introduced here showed an example in the north Atlantic where monthly
use daily mean meteorological values and/or previous- mean SST was higher by about 0.2 K in the case that di-
day information, it may be difficult to utilize these urnal SST variations were included than in the case that
schemes for discussing simultaneous air-sea interaction they were ignored. This difference reduces the net heat
processes in a coupled model. However, they will be use- flux of 5 W m2 entering the ocean.
ful to test the sensitivity of the atmosphere to the diurnal Fairall et al. (1996) showed that the cool skin de-
variation of SST. creased the net heat flux from the ocean by about 11
W m2 and the warm layer increased it by about 4 W m2
5. Effect of Diurnal Thermocline on Air-Sea Flux on average over 70 days during COARE. The effect of
Estimations the warm layer may seem to be fairly small, but this was
the mean value including the cases when the warm layer
5.1 Air-sea heat flux estimation did not develop. Fairall et al. (1996) and Ward (2006)
Sverdrup et al. (1942) suggested the importance of indicated that the net heat flux from the ocean can in-
the diurnal SST variability on air-sea heat exchange. Since crease 5060 W m2 in the daytime under calm and clear
the atmosphere contacts the sea skin, not the water at a conditions due to the effect of the warm layer (Fig. 3).
few meters depth, evaluating the temperature difference Clayson and Curry (1996) estimated surface turbulent heat
across the warm layer and skin layer is indispensable for fluxes during COARE from satellite data and compared
accurate air-sea heat flux estimation. An error of 1 K in them with in situ measurements. They showed that deter-
SSTskin can lead to an error of 27 W m 2 in net surface mining the fluxes every 3 h from interpolated satellite-
heat flux in the tropical western Pacific (Webster et al., derived input variables, i.e., including diurnal cycles in
1996). Furthermore, Cornillon and Stramma (1985) SSTskin and atmospheric variables, improved the estimate

730 Y. Kawai and A. Wada


Fig. 8. Diurnal amplitude of SSTskin estimated by Webster et al.s (1996) empirical model as a function of daily peak solar
radiation for different values of daily mean wind speed and precipitation. Solid, broken, and dotted lines represent 0, 1, and 5
mm h1 of daily mean precipitation rate, respectively. Black, magenta, red, blue, and green lines represent 0.1, 1, 3, 5, and 10
m s1 of daily mean wind speed, respectively.

of daily-mean surface fluxes. However, they did not evalu- this scheme increased the latent heat flux during the day-
ate the effect of the diurnal variation of SSTskin on the time by 1020 W m2 and reduced it in the nighttime by
fluxes. 05 W m2. The increase in time-mean net heat loss of
Zeng and Dickinson (1998) investigated the impact the ocean was about 10 W m2. Zeng and Beljaars (2005)
of the diurnal variation of SSTskin on surface fluxes over reported that incorporating their SSTskin scheme (see Sub-
the equatorial Pacific using TOGA Tropical Atmosphere section 4.2) into an operational forecasting model changed
Ocean (TAO) buoy hourly data from 95W to 137E in ensemble annual mean surface latent heat flux by more
19901996. The surface latent and sensible heat fluxes than 10 W m2 over several regions in the north Atlantic.
showed clear diurnal variability, and the average diurnal The authors also checked the impact of the warm
amplitudes of the latent and sensible heat fluxes were 19.7 layer on the air-sea heat transfer by a simple numerical
and 5.6 W m2, respectively. They also calculated the heat experiment. The test data used here were obtained with a
fluxes by replacing hourly SSTskin with daily or monthly moored buoy of the Triangle Trans-Ocean buoy Network
mean SSTskin. Figure 9 shows the examples of the differ- (TRITON) in the western tropical Pacific at 2.07N,
ences between the heat fluxes with the hourly SSTskin and 138.06E during 313 March 2004. The authors simu-
those with the daily or monthly SSTskin. Evidently the lated near-surface temperature using the buoy-observed
diurnal cycle in SSTskin is the main cause of the diurnal meteorological data and Kawai and Kawamuras (2000)
variability of latent and sensible heat fluxes. They sug- one-dimensional model. The details of the observations
gested that numerical modeling may require the inclu- and the model simulation are reported in Kawai et al.
sion of the diurnal SSTskin variation. Parsons et al. (2000) (2006b). The diurnal temperature variation was large on
also showed that the average diurnal amplitudes of sur- 3, 7 and 9 March, and the model could reproduce the vari-
face sensible and latent heat fluxes were about 4 and 35 ation of SST0.3m approximately (Kawai et al., 2006b). This
W m2, respectively, within the inner intensive flux array simulation result of SSTskin is shown in Fig. 10(a) with a
of the COARE experiment in mid-November 1992. solid line, and is called the control run. The layer be-
Schiller and Godfrey (2005) examined the effect of tween the sea surface and 1.5-m depth was then forcibly
the diurnal variability of SST on surface latent heat flux mixed by setting the minimum of the eddy diffusion co-
during COARE using a one-dimensional coupled model efficients to 5.0 10 4 m2s1 above 1.5-m depth. This
with their sublayer scheme (see Subsection 4.2). Use of simulation is referred to as the no-warm-layer run (bro-

Diurnal SST Variation and Its Impact 731


ken line in Fig. 10). In the no-warm-layer run the diurnal
thermocline was destroyed and the temperature became
vertically homogeneous within 01.5 m depth, even in
the daytime (Fig. 10(b)). The diurnal amplitude of SSTskin
decreased by about 1.5 K in the no-warm-layer run in the
daytime of 7 and 9 March. The no-warm-layer run al-
most corresponds to the case that a model applies a coarse
vertical grid of 1.5 m, which cannot resolve the shallow
warm layer.
If the warm layer cannot be reproduced correctly,
the surface net heat flux from the ocean decreases by 20
40 W m2 in the daytime (Fig. 11). The integrated differ-
ence of the net heat flux through a day exceeds 0.9
MJ m 2 on 7 and 9 March (Table 3). About 64% of the
total heat difference is due to latent heat. The effect of
the warm layer is to retain the incident solar energy near
the sea surface and increase the heat transfer back to the
atmosphere, mainly as latent heat. The heat of 0.9
MJ m 2, which can raise the temperature of the 10-m-
thick water column by only 0.02 K, may not be important
for the ocean, but it will not be negligible for the atmos-
phere because the thermal capacity of the air is much
smaller and such an amount of excessive heat from the Fig. 9. Mean diurnal cycles of surface flux differences at 0N,
ocean can destabilize the lower atmosphere. 156E (ab), and 2S, 95W (cd). Latent (a and c) and
Recently several kinds of global surface flux datasets sensible (b and d) heat flux differences using hourly versus
daily (monthly) mean SSTskin are denoted by solid (dotted)
have been developed using satellite and/or reanalysis data
lines. Abscissa represents local time. From Zeng and
(cf. Kubota et al., 2003; Curry et al., 2004). In some of Dickinson (1998), Copyright 1998 American Geophysical
these datasets, the SST produced by optimum interpola- Union. Reproduced by permission of American Geophysi-
tion (e.g., Reynolds and Smith, 1994; Reynolds et al., cal Union.
2002) is used for the flux calculation. However, such SST
data lack diurnal and day-to-day variability. Yu et al.
(2004) suggested that the absence of the high-frequency
variations in SST appeared to cause the degradation in perature dependence of 4.23% K1 for constant total CO 2
the accuracy of air-sea temperature and specific humid- was similar to the observed fCO2 (Fig. 12). This means
ity differences. Consideration of the high-frequency that fCO2 near the surface can vary widely in association
SSTskin variations will be necessary to improve the qual- with diurnal warming. The diurnal amplitude of the cal-
ity of surface flux data. culated fCO2 was 1020 atm.
Ward et al. (2004a) evaluated the effects of the skin
5.2 Air-sea gas flux estimation layer and the warm layer on the air-sea CO2 exchange in
The solubility of gases in seawater depends on tem- the eastern equatorial Pacific during the GasEx-2001
perature. Hence diurnal increases in near-surface tempera- cruise. Basically, the skin effect consistently suppresses
ture can change the concentrations of CO2 and O2 and the CO2 emission from the oceans through a day, while
modify the air-sea gas flux. Researchers working on the the formation of the warm layer increases it only in the
air-sea gas exchange have been paying attention to the daytime. According to their evaluation, the mean decrease
diurnal warming and the skin effect. For example, of the surface CO2 flux due to the skin effect was about
Soloviev et al. (2001) reported an observed example that 2% with a maximum of almost 4%. The increase of the
dissolved oxygen concentration just near the surface (shal- CO2 flux due to the warm layer could exceed 6% tempo-
lower than 0.1-m depth) was lower by about 5 ml l1 than rarily, although the mean value throughout the observa-
that at 0.5-m depth or deeper in association with the for- tion period was only about 0.7%. They implied that the
mation of the diurnal thermocline. McNeil and Merlivat effect of the warm layer may be less important for the
(1996) revealed the diurnal variations of water tempera- air-sea CO2 exchange than the skin effect, because while
ture and dissolved CO2 fugacity (fCO2) observed at 2-m the warm layer forms only in the daytime of a calm and
depth in the Mediterranean Sea, and indicated that the clear day, the skin layer exists even in windy conditions
fCO2 predicted from observed SST2m and an fCO2-tem- (Donlon et al., 2002). The relative increase due to the

732 Y. Kawai and A. Wada


Fig. 11. Difference of net air-sea heat flux between the two
runs (control run minus no-warm-layer run, upward is posi-
tive) at 2.07N, 138.06E in March 2004. Shortwave radia-
tion is not included.

6. Impact on the Atmosphere


The diurnal variation of SST is strongly affected by
meteorological conditions. Can the diurnal SST variation
really significantly affect the properties of the atmos-
phere? As yet, we do not know the impact of the diurnal
SST variation on physical properties or processes in the
Fig. 10. (a) Time series of the model-simulated SSTskin in the atmosphere, especially in the boundary layer. Large diur-
no-warm-layer run (broken line), and that in the control run nal SST variations must be accompanied by a weak wind,
(solid line) at 2.07N, 138.06E in March 2004. (b) Their which suppresses the turbulent heat transfer from the
vertical temperature profiles at 1600 LST on 9 March. ocean to the atmosphere. Zhang (2005) pointed out that
the atmosphere does not see SST; it only senses it through
surface fluxes. From this viewpoint, paradoxically, the
atmosphere near the surface may sense that the sea sur-
warm layer effect shown by Ward et al. (2004a) was cer- face is cool in calm and clear conditions. Chen and Houze
tainly small, but this can correspond to a large absolute (1997) indicated that the diurnal cycle in the surface air
increase in the regions such as the eastern tropical Pa- temperature was not completely dependent on that in
cific where the absolute value of the CO2 flux is large. SSTskin, although they suggested the importance of the
They also indicated that the estimated difference of fCO2 diurnal heating of the sea surface on cloud convection
across the warm layer of a few meters thickness can reach over the Pacific warm pool.
20 atm or more (Fig. 13). In the eastern equatorial Pa- On the other hand, as mentioned in Subsection 5.1 it
cific, the diurnal SST amplitude becomes larger during has been reported that the diurnal SST rise in the equato-
La Nias and lower during El Nios. Hence it is expected rial Pacific can increase the surface net heat flux by the
that the impact of the warm layer on the surface CO2 flux order of 10 Wm2, which is probably non-negligible for
will vary with the ENSO cycle (Cronin and Kessler, 2002). atmospheric physical processes. Several researchers as-
The skin effect on the global CO2 budget has also sume that the diurnal SST variability may have some
been assessed by several researchers (e.g., Sarmiento and impact on the atmosphere. For example, Dai and Trenberth
Sundquist, 1992; Robertson and Watson, 1992; Van Scoy (2004), using a fully coupled climate system model with-
et al., 1995; Wong et al., 1995), but this issue is beyond out the diurnal SST variation, indicated that simulated
the scope of this paper, and we do not review it here. The diurnal cycles in surface air temperature, pressure and
studies mentioned in this subsection indicated that a care- precipitation over the oceans were much weaker than the
ful consideration of the ocean surface boundary layer and observed values. They inferred that the lack of the diur-
its high-frequency variability will be significant in stud- nal cycle in SST was a significant deficiency. The diur-
ies of the global material circulation and climate change. nal variability of SST in the tropics is basically large,

Diurnal SST Variation and Its Impact 733


Table 3. Daily-integrated differences of the surface heat fluxes
between the control run and the no-warm-layer run. Unit is
MJ m 2.

3 March 7 March 9 March

Net 0.649 0.912 0.990


Sensible heat 0.080 0.112 0.117
Latent heat 0.417 0.590 0.631
Longwave radition 0.152 0.210 0.242

and has been intensively investigated since COARE. In


this section we first review studies and hypotheses on the
effect of the diurnal SST variation on the tropical atmos-
phere. Other topics in the mid and high latitudes are then Fig. 12. Observations during 5 days in August 1995 of 2-m-
summarized. depth water temperature (T, dashed line) and fugacity of
CO2 (fCO2, bold solid line) at a mooring site between Nice
6.1 Tropics and Corsica in the Mediterranean Sea (43.42N, 7.87E).
The diurnal variability of cloud convection over the Thin solid line is the predicted fCO2 from the observed T
ocean has been observed, although its mechanism has not and an fCO2-temperature relation of 4.23% K1 for constant
yet been completely solved. In general, the diurnal varia- total CO2. From McNeil and Merlivat (1996), Copyright
tions of cumulus convections over the ocean and land are 1996 American Geophysical Union. Reproduced by permis-
different from each other. The convective activity is en- sion of American Geophysical Union.
hanced in the late afternoon to evening over continents
and large islands, while it attains its maximum in the early
morning over the ocean (e.g., Gray and Jacobson, 1977;
Janowiak et al., 1994; Yang and Slingo, 2001; Dai and ity. Parsons et al. (2000) and Chen and Houze (1997)
Trenberth, 2004). Various kinds of mechanisms have been claimed that the absorption of shortwave radiation in the
proposed for the formation of diurnal deep convection atmospheric boundary layer will be as important (or more
over the ocean: the direct interaction between radiation important) in the diurnal variation in CIN as (than) the
and convection (Randall et al., 1991); the effect of the diurnal cycle in SST and the resulting changes in the
horizontal distribution of clouds on radiation (Gray and fluxes. Slingo et al. (2003) suggested the possibility that
Jacobson, 1977); the radiative interaction between sur- the diurnal SST rise may act as a trigger for the shallow
face and clouds (Chen and Houze, 1997); the diurnal vari- cloud convection. The basic paradigm proposed by them
ation of available precipitable water due to the diurnal is as follows: the diurnal cycle in SST leads to a trigger-
radiative cooling/heating cycle (Sui et al., 1997a); and ing of convection in the inactive phase of MJO. In turn
gravity wave forcing induced by the nearby continental the cumulus congestus clouds gradually moisten the free
diurnal cycle of convection (Mapes et al., 2003). troposphere and prepare a favorable condition for deep
Analyses of the COARE observation results revealed convection. This preconditioning may set the time scale
a more detailed aspect of the diurnal variation of convec- for the following active phase of MJO. These hypotheses
tion over the warm pool. During the convectively active have not yet been corroborated.
phase of MJO, the large, deep convective systems tended Li et al. (2001) performed numerical experiments
to reach a maximum before dawn, as mentioned above. using a global atmospheric model with or without the di-
On the other hand, during the convectively suppressed urnal variation of SSTskin. They indicated that the phases
phase of MJO, when the diurnal SST variation was large, of the intraseasonal variations of surface flux and pre-
shallow precipitating clouds were most abundant (Fig. cipitation simulated with diurnal SSTskin variations were
14). Many of these shallow clouds occurred in the after- closer to those of the observations in the warm pool dur-
noon near the time of maximum SST, unlike the diurnal ing COARE, compared to those simulated without the
variation of the deep convection in the active phase (e.g., diurnal SSTskin variations. The diurnal SSTskin variation
Chen and Houze, 1997; Sui et al., 1997a; Johnson et al., brought the increase of precipitation especially over the
1999). Parsons et al. (2000) showed from COARE sound- tropical Indian Ocean and the western tropical Pacific,
ing data that there was a detectable diurnal cycle in con- although the ratio of the increase to the mean value was
vective inhibition (CIN), with a minimum in the late af- not more than 10%.
ternoon just preceding the maximum in convective activ- Woolnough et al. (2007) also examined the impact

734 Y. Kawai and A. Wada


Clayson and Chen (2002) investigated the sensitiv-
ity of the atmosphere to SST as the lower boundary con-
dition in the tropical Pacific using a coupled single-col-
umn model. They showed that use of the simulated SSTskin
as the interfacial temperature rather than SSTdepth resulted
in large differences in the atmospheric profiles of tem-
perature, moisture, and cloud amount. Figure 15 shows
the cloud amount simulated with the SSTskin (baseline
simulation), and that with the SST4.5m (4.5-m tempera-
ture simulation). The SST4.5m was higher by 0.25 K than
the SSTskin on average due to the skin effect, and showed
less diurnal variability. The baseline simulation has lower
low-level cloud amounts during periods 1 and 3, and
higher amounts during period 2, compared with the 4.5-
m temperature simulation. The cloud amount differences
in the low level are larger than that in the mid level. One
of their interesting indications is that both the skin effect
and the diurnal variability are important for the atmos-
phere, and which of the effects is dominant varies with
the periods. Clayson and Chen (2002) also indicated that
the model atmosphere was sensitive to the scheme of sur-
face turbulent heat fluxes.
Deser and Smith (1998) indicated from TAO buoy
data that the mean diurnal amplitude of SST showed a
local maximum over the cold tongue in the eastern equa-
torial Pacific. They proposed a hypothesis that this large
diurnal SST variation over the cold tongue may affect
the zonally symmetric diurnal cycle of equatorial wind
divergence.
Fig. 13. Series of SkinDeEP temperature profiles acquired on
8 Feb. 2001 in the eastern equatorial Pacific (GasEx-2001) 6.2 Mid and high latitudes
and the extrapolation of the aqueous fCO 2 to the surface Flament et al. (1994) found streaks of large diurnal
from an fCO2-temperature dependence of 4.23% K1 for sea surface warming that occurred in spring off Califor-
constant total CO2. The local times are indicated for each nia from satellite data. These streaks were at least 50 km
plot. From Ward et al. (2004a), Copyright 2004 American
long and 48 km wide. They hypothesized that these
Geophysical Union. Reproduced by permission of Ameri-
streaks may interact with roll-like circulations in the at-
can Geophysical Union.
mospheric boundary layer, and proposed two kinds of
hypotheses to explain this phenomenon. These hypoth-
eses are interesting, but nobody has verified them yet.
of the diurnal mixing of the upper ocean on MJO by cou- Kawai et al. (2006a) investigated the impact of diur-
pled model experiments. They compared the result of nal sea surface warming on a local atmospheric circula-
constant-SST simulation with the result simulated by cou- tion over Mutsu Bay, which is located at the northern end
pling to a full dynamical ocean model of 10-m vertical of Honshu Island in Japan around 41N, 141E, on a clear
resolution in the upper ocean, or that by coupling to a summer day. Over this region small sea breeze circula-
one-dimensional ocean mixed layer model of 1-m reso- tions and mountain up-slope wind circulations are com-
lution. This 1-m resolution mixed layer model could re- bined during the daytime, and a unique atmospheric cir-
produce large diurnal and intraseasonal SST variations. culation is formed. The result of their simple model ex-
The experiment with the mixed layer model showed the periment showed that the SSTskin rise in the daytime weak-
best improvement in prediction skill, especially for the ens this atmospheric circulation due to the decrease of
phase of the MJO over the Indian Ocean and the western the land-sea temperature difference. This results in the
Pacific. They indicated the role of the ocean in determin- increase in surface air temperature over coastal areas. The
ing the propagation characteristics of the MJO, and the effect of the diurnal sea surface warming on the sea breeze
significance of the representation of the diurnal cycle in circulation is expected to obtain over other coastal areas
the upper ocean. throughout the world, especially in the tropics. Yang and

Diurnal SST Variation and Its Impact 735


became large in the Bay of Bengal and the South China
Sea during the transition period of the Asian monsoon in
boreal spring. It was also reported from in situ observa-
tions that the diurnal SST cycle in the South China Sea
was dominant before the onset of the monsoon, and de-
creased thereafter (D. Wang, the presentation in the 11th
Ocean Observations Panel for Climate, 2006). A suggested
possibility was that the diurnal variability of SST may
affect the monsoon onset, but no corroboration has yet
been obtained.
Kawai and Kawamura (2005) indicated that the di-
urnal amplitude of SST also became large, even in the
Sea of Okhotsk from spring to summer. Low-level clouds
and fog frequently cover this region in summer. Tachibana
et al. (2004) pointed out that the appearance and disap-
pearance of fog in the Sea of Okhotsk will be controlled
by air-sea interaction. They reported from ship observa-
tions that when warm air was advected in the lowest level,
the air cooled while the upper ocean warmed. In this case
the atmospheric mixed layer could not develop and fog
did not occur, which led to strong insolation at the sea
surface. Therefore, thermal stratification was formed
within a few meters depth. The large diurnal rise of SST
may, in turn, change the stability of the lowest atmos-
phere. Furthermore, the formation of the shallow diurnal
stratification may affect the variations of SST and the
ocean mixed layer on a longer time scale (see Subsection
Fig. 14. Daily-average number of low or cumulus (04 km), 7.1).
middle or congestus (59 km), and high or cumulonimbus
(1116 km) radio-echo tops for a cruise of R/V Vickers dur-
ing COARE, and SST observed with the Improved Mete-
7. Issues in Modeling and Observation of Diurnal
orological instrument (IMET) buoy at 1.75S, 156E. In the SST Variation
upper three panels, solid curve segments refer to periods
when convective echoes organized on the Mesoscale Con- 7.1 Numerical modeling
vective System (MCS) scale (>100 km) and dotted segments a. Surface forcing
to periods when only sub-MCS or isolated cells existed. In order to reproduce realistic diurnal variation of
Reproduced from Johnson et al. (1999). SST using an ocean model, accurate surface forcing is
required as a boundary condition. In particular, insola-
tion with a diurnal cycle is important in reproducing the
diurnal variation. Recently many model researchers have
Slingo (2001) indicated that the strong diurnal signals of investigated the effect of diurnal variability of surface
convection over land were spread out many hundreds of forcing on modeling of SST and ocean mixed layer (e.g.,
kilometers away over the Bay of Bengal and the coastal Sui et al., 1997b; Shinoda and Hendon, 1998; McCreary
seas in the Maritime Continent. They suggested that this et al., 2001; Schiller and Godfrey, 2003; Bernie et al.,
phenomenon is affected by complex land-sea and moun- 2005; Lee and Liu, 2005; Shinoda, 2005; Danabasoglu et
tain-valley breezes. The diurnal cycle in SST may affect al., 2006). These studies indicate that the diurnal varia-
this phenomenon through modification of the land-sea tion of surface forcing, especially that of insolation, plays
breeze circulation. Yang and Slingo (2001) and Slingo et an important role in reproducing diurnal and intraseasonal
al. (2003) point to the need for a proper parameterization variations, or a mean state in the upper ocean. The diur-
of the effect of land-sea breezes in a numerical model for nal change of the mixing process of the upper layer is not
better simulation of the diurnal cycle in convection. The only related to the diurnal SST variability, but is also es-
effect of the diurnal SST variation should also be included sential for the intraseasonal SST variation. The inclusion
in the parameterization or model. of the diurnal cycle of insolation enables one to simulate
Stuart-Menteth et al. (2003) and Kawai and a more realistic temperature and depth of mixed layer.
Kawamura (2005) showed that the diurnal SST variation The temporal mean of tropical SST simulated with the

736 Y. Kawai and A. Wada


Fig. 15. Time series of (a) daily-averaged mid-level (400700 hPa) cloud amount, and (b) low-level (below 700 hPa) cloud
amount for the baseline simulation (solid line) and the 4.5-m temperature simulation (dashed line) with a coupled atmos-
phere-ocean single-column model during COARE. Reproduced from Clayson and Chen (2002).

diurnal cycle of insolation tends to be higher than with- tropics cannot be reproduced if an inappropriate
out it. parameterization for shortwave radiation is used.
Figure 16 shows the result of a numerical experi- An approximate formula of a polynomial exponen-
ment performed by Bernie et al. (2005), who simulated tial function (e.g., Paulson and Simpson, 1977) consider-
mixed layer depth and temperature with hourly surface ing only the simple classification of water type (e.g.,
fluxes and with daily mean fluxes, respectively. The am- Jerlov, 1968) is convenient and is often used to calculate
plitude of the intraseasonal SST variation clearly de- the downward shortwave radiation flux in seawater. How-
creases if the diurnal variations of the surface fluxes are ever, this parameterization is not always satisfactory (e.g.,
neglected. The mixed-layer depth simulated with the Sui et al., 1997b), and cannot consider the interaction
hourly fluxes becomes a little deeper than that with the between physical and biological processes. Recently a
daily fluxes. Sui et al. (1997b) insisted that, because of significantly improved parameterization that depends on
the asymmetric heating rate of the diurnal cycle, the vari- upper-ocean chlorophyll-a concentration, cloud amount,
ation of mixed-layer properties on the diurnal timescale and solar zenith angle has been proposed (Ohlmann and
is nonlinearly related to the intraseasonal variability. The Siegel, 2000). Fairall et al. (1996) originally adopted
diurnal change of the ocean mixed layer will affect the shortwave-radiation parameterizations that depended on
phase of intraseasonal atmospheric variation through the only depth for their skin-layer and warm-layer models.
amplification of the intraseasonal SST variation (cf. Li et Ohlmann and Siegel (2000) and Wick et al. (2005) re-
al., 2001; Woolnough et al., 2007). The inclusion of the placed them with Ohlmann and Siegels parameterization,
mixed-layer deepening/shoaling process on the diurnal and compared the temperature and surface fluxes simu-
scale cannot be neglected in biological modeling lated by this modified version of Fairall et al.s model
(McCreary et al., 2001). with those by the original model. If Ohlmann and Siegels
b. Solar extinction modeling parameterization is adopted, due to the reduction of the
Unlike longwave radiation, sensible and latent heat absorption of insolation, the skin effect is strengthened
fluxes, shortwave radiation can penetrate the ocean. De- and the temperature difference across the warm layer is
termination of vertical distribution of warming by decreased. As a result, SSTskin and the net heat flux from
shortwave radiation in the upper ocean is critical to re- the ocean can be reduced by about 0.2 K and 5 Wm1
producing accurate diurnal and intraseasonal variations under calm and clear conditions, respectively.
of SST (e.g., Kantha and Clayson, 1994; Sui et al., 1997b; As reviewed here, use of an improved
Shinoda, 2005; Wick et al., 2005; Ward, 2006; Clayson parameterization for shortwave radiation in the ocean is
and Weitlich, 2007). Sui et al. (1997b) pointed out that important to simulate SST and surface fluxes in a numeri-
realistic diurnal and intraseasonal SST variations in the cal model accurately. It is known that biological proc-

Diurnal SST Variation and Its Impact 737


esses such as increase of phytoplankton and cyanobacteria
can significantly affect SST (e.g., Kahru et al., 1993).
Siegel et al. (1995) found that in the warm pool during
COARE, the penetration depth of shortwave radiation
decreased substantially after westerly bursts. This was due
to the upward mixing of nutrients and subsequent
phytoplankton growth. Godfrey et al. (1998) suggested
that plankton dynamics may have a significant influence
on SST in the warm pool too. It has also been reported
that the chlorophyll-a concentration in the upper ocean
significantly increases in response to tropical cyclones
(e.g., Subrahmanyam et al., 2002; Siswanto et al., 2007).
The formula proposed by Ohlmann and Siegel will be
useful when considering the physical-biological interac-
tion in relation to air-sea interaction.
c. Time step and vertical resolution
Restricted computational resources make it suffi- Fig. 16. (a) Sample SST and (b) turbulent boundary layer depth
ciently difficult to refine the temporal and vertical reso- time series from the control integration with hourly fluxes
lution of an OGCM so that a sharp diurnal thermocline (Run CTL, solid line), and the sensitivity experiment with
can be reproduced. Although even a model with a coarse daily mean fluxes (Run 24HR, dotted line). IMET buoy data
vertical resolution of 1015 m in the upper ocean can at 2S, 156E during COARE and a one-dimensional mixed
simulate the diurnal variation to a certain degree (Schiller layer model were used for the simulations. Dashed line
and Godfrey, 2003; Danabasoglu et al., 2006), more real- shows the daily mean SST from the control integration to
emphasize the intraseasonal variability. Reproduced from
istic simulations need a much finer vertical resolution in
Bernie et al. (2005).
the upper ocean. Bernie et al. (2005) discussed the tem-
poral resolution of forcing flux and the vertical resolu-
tion required to accurately reproduce the diurnal varia-
tion of SST by a numerical model. They indicated that a 7.2 Observations
temporal resolution of less than 3 h and a vertical resolu- Satellite observations, especially by such microwave
tion of less than 1 m should be specified in order to re- sensors as TMI and AMSR-E, are very effective, indeed
produce more than 90% of the amplitude of the diurnal indispensable for research into high-frequency SST vari-
SST variation. ation. In order to resolve the global diurnal cycle, flights
d. Other model parameterizations of at least two satellites with microwave sensors are de-
Modeling methodology remains affected by other sirable. High-quality, sustainable satellite observations
problems, one of which is uncertainty in the bulk coeffi- must be continued.
cients used to estimate air-sea fluxes (e.g., Clayson and However, uncertainty in satellite-derived SST is una-
Chen, 2002). Another is concerned with schemes of an voidable. High-quality, continuous in situ observations
atmospheric model. A coupled GCM (CGCM) with a short of near-surface temperature are also necessary in order
coupling interval of 13 h between an atmospheric GCM to minimize satellite SST errors and to determine the ver-
and an OGCM can produce diurnal ocean variations tical temperature profile. At present moored buoys are
(Danabasoglu et al., 2006). However, even if correct di- densely arranged only in the tropics and some coastal ar-
urnal SSTskin variations are supplied to an atmospheric eas. As mentioned above, the diurnal variability of SST
model as a lower boundary condition, the model atmos- will be also important in the extratropics, and we need a
phere cannot correctly respond to the diurnal variations dense in situ observation system to observe the diurnal
of SSTskin without appropriate parameterizations. Present change of the upper ocean even in the mid- and high-
CGCMs still have deficiencies in the dynamics for low- latitudes.
level atmospheric convergence and in their physical In situ SST measurement also has some problems.
parameterizations for the planetary boundary layer, cloud For ship SST data, it is well known that the SSTs reported
and precipitation formation, moist convectionespecially from voluntary observing ships are noisier than buoy
at its initiation, as well as the deficiency that simulated SSTs, and can have a warm bias due to the heating in the
diurnal SST variations are too small (Dai and Trenberth, engine room (e.g., Saur, 1963; Reynolds and Smith, 1994;
2004). These atmospheric parameterizations also need to Emery et al., 2001; Reynolds et al., 2002). The SSTs ob-
be improved in order to study air-sea interaction on a di- served with buoys are usually considered to be more reli-
urnal time scale. able, and are utilized for the tuning and validation of sat-

738 Y. Kawai and A. Wada


ellite SSTs. However, Kawai and Kawamura (2000) and SST variation can be approximately simulated by the
Kawai et al. (2006b) indicated that the temperature ob- models, although there still remain some challenges in
served with a moored buoy at about 1-m depth would precise modeling.
deviate from the actual temperature at the nominal depth The atmosphere feels SSTskin, not SSTdepth, through
when a sharp diurnal thermocline develops just near the surface heat fluxes. Hence the estimations of air-sea heat
surface. Although the exact reason for this still remains and gas fluxes are susceptible to the diurnal SST vari-
unclear, the turbulence and/or heating induced by a buoy ability and the skin effect. Previous studies of the surface
hull are suspected. flux estimations were then summarized (Section 5). While
A careful temperature measurement technique that the cool skin layer almost always exists through a day,
avoids disturbing the near-surface temperature field is re- even in windy conditions, the warm layer is formed only
quired when a sharp diurnal thermocline is formed. One in the daytime of a calm and clear day. Therefore, the
of the solutions is use of a compact profiling float. Ward skin effect may be more important for the flux estimation
and Minnett (2001) and Ward et al. (2004b) developed than the effect of the warm layer. However, for the heat
the SkinDeEP autonomous profiler, which can measure flux, it is indicated that the consideration of the warm
near-surface temperature with a fine vertical resolution layer can increase the net heat flux by 5060 W m2 in
without disturbance. SkinDeEP is similar to an Argo-type the daytime, and its temporal mean exceeds 10 W m 2.
float, but it is smaller and its temperature sensors pro- Furthermore, the significant effect of the skin layer and
trude several tens of centimeters from the top of the body the warm layer on air-sea gas exchange has already been
to avoid distorting the stratification. Although this float pointed out, and the surface mixing process on the diur-
is excellent in near-surface temperature measurement, nal scale can also affect marine biology (McCreary et al.,
only a few prototypes have been built and they are not 2001). Accurate physical-chemical-biological modeling
equipped with telecommunication functionality. At and material circulation studies will need knowledge of
present, many profiling floats are deployed globally un- the oceans diurnal variability.
der the international cooperation of the Argo Project (e.g., Recent studies of the relation between the diurnal
Argo Science Team, 2001; Argo Project Office, 2006), SST variation and air-sea interaction were then introduced
but the normal profiling floats do not measure the sur- (Section 6). Whether the diurnal SST variation really af-
face layer above about 5-m depth in order to prevent deg- fects the atmospheric physics significantly is a difficult
radation of sensor performance caused by contaminants but interesting question. A few studies have showed that
near the sea surface, such as oil (e.g., Kobayashi et al., the diurnal variation and/or the skin effect can modify
2004; Riser and Wijffels, 2005). If the profiling floats precipitation and surface fluxes variations on an
are improved to measure near-surface temperature with intraseasonal scale, the vertical profile of the air tempera-
fine vertical and temporal resolution, this will become a ture and humidity, and a sea-breeze circulation. Although
powerful tool for studies on the diurnal SST variability. at present the investigation of this subject is not yet ad-
equate, many researchers expect that the diurnal cycle in
8. Summary SST will have some impact on the atmosphere, and an
The importance of the diurnal variability of SST is attempt to simulate the diurnal variation of the upper
becoming increasingly clearly recognized. This paper has ocean in a GCM has already begun. The necessity of this
summarized studies of the diurnal SST variation and its attempt at a numerical model has also summarized (Sub-
possible impacts on air-sea interaction. We first introduced section 7.1). Fine vertical/temporal resolution and appro-
the latest definitions of the several kinds of SST (Sec- priate parameterization for shortwave extinction in the
tion 2), and reviewed the observational facts about the ocean are required to accurately simulate large diurnal
diurnal SST variation (Section 3). Basically, when inso- SST rise. Some schemes that enable us to reproduce real-
lation is strong and wind speed is low, the diurnal rise of istic diurnal SST variations without enormous computa-
SST becomes large. Early studies indicated that the diur- tional load have been proposed recently (e.g., Schiller and
nal amplitude of SSTdepth was 0.20.6 K in the low and Godfrey, 2005; Zeng and Beljaars, 2005).
mid latitudes on average, reaching about 1.5 K under clear In conclusion, the issues to be clarified can be sum-
and calm conditions. Studies conducted during the past marized as follows:
two decades have revealed from in situ and satellite ob- (1) A few studies have pointed out from numerical
servations that the diurnal rise of SSTskin or SSTsubskin can experiments that the diurnal SST variation and/or the skin
reach 5 K or more in extreme cases. Satellite observation effect can significantly affect the atmospheric field over
also clarified that the large diurnal rise can occur over the warm pool (e.g., Li et al., 2001; Clayson and Chen,
wide areas in a specific season. Various kinds of model 2002). However, the processes through which the differ-
have been proposed and utilized to simulate diurnal vari- ences in cloudiness and humidity were caused were not
ations in the upper ocean (Section 4). The large diurnal discussed, and still remain unknown.

Diurnal SST Variation and Its Impact 739


(2) The diurnal temperature rise and the tempera- Acknowledgements
ture drop across the skin layer increase or decrease the The authors would like to greatly thank C. J. Donlon,
air-sea sensible heat flux, and thus will affect the heat H. Kawamura, G. A. Wick and all the members of The
content and height of the atmospheric boundary layer. The GHRSST-PP Science Team, K. Yoneyama, K. Ando, M.
perturbation of only a few W m2 of the sensible heat flux, Katsumata, T. Kobayashi of JAMSTEC, J. Ishizaka of
which will be negligible for the ocean, may significantly Nagasaki University, and T. Shimada and H. Qin of
affect the lower atmosphere. The changes of the atmos- Tohoku University for providing us with useful informa-
pheric mixed layer and the surface wind field related with tion and support. We also appreciate valuable comments
the diurnal SST variation need to be investigated inten- of two anonymous reviewers. The AMSR-E data used in
sively. Furthermore, whether large diurnal SST rise re- this paper are produced by Remote Sensing Systems and
ally can be the trigger for the shallow cumulus convec- sponsored by the NASA Earth Science REASoN
tion (Slingo et al., 2003), which may control the precon- DISCOVER Project and the AMSR-E Science Team. The
ditioning of the MJO active phase, has not been confirmed data are available at [www.remss.com].
yet, either.
(3) The effect of the diurnal SST rise on a coastal References
local atmospheric circulation has been indicated by Kawai Argo Project Office (2006): Argonautics, Newsletter of the In-
et al. (2006a). They reported only an idealized calm and ternational Argo Project, No. 7. Argo Project Office, Uni-
clear case. More detailed numerical experiments are nec- versity of California, San Diego, 12 pp. (available at: http:
//www.argo.ucsd.edu/Frnewsletter.html).
essary for more realistic conditions and/or different ar-
Argo Science Team (2001): Argo: The global array of profiling
eas, especially for the tropics. Consideration of this ef-
floats. p. 248258. In Observing the Oceans in the 21st
fect may be significant for practical weather forecasting Century, ed. by C. J. Koblinski and N. R. Smith, GODAE
in coastal regions. Project Office, Bureau of Meteorology, Melbourne, Aus-
(4) While many researchers interested in the diur- tralia.
nal SST variation pay much attention to the tropics and Barton, I. J., P. J. Minnett, K. A. Maillet, C. J. Donlon, S. J.
subtropics, the diurnal variation in the high latitudes has Hook, A. T. Jessup and T. J. Nightingale (2004): The Mi-
rarely been noted. However, as shown in Fig. 5, the diur- ami2001 infrared radiometer calibration and
nal SST rise can become large even in the high latitudes intercomparison. Part II: Shipboard results. J. Atmos. Oce-
in summer. In situ observations that can resolve the diur- anic Technol., 21, 268283.
nal SST cycle are not sufficiently close to the polar re- Bernie, D. J., S. J. Woolnough, J. M. Slingo and E. Guilyardi
(2005): Modeling of diurnal and intraseasonal variability
gions. Intensive studies focusing on the high latitudes
of the ocean mixed layer. J. Climate, 18, 11901202.
should be conducted.
Bruce, J. G. and E. Firing (1974). Temperature measurements
(5) As mentioned in Subsection 7.1.d, some in the upper 10 m with modified expendable
parameterizations in an atmospheric model would not be bathythermograph Probes. J. Geophys. Res., 79, 41104111.
appropriate for simulating atmospheric response to the Chen, S. S. and R. A. Houze, Jr. (1997): Diurnal variation and
diurnal SST variation. Further improvement of the model life-cycle of deep convective systems over the tropical Pa-
parameterizations will be necessary. cific warm pool. Q. J. R. Meteorol. Soc., 123, 357388.
(6) When a sharp and shallow diurnal thermocline Clayson, C. A. and A. Chen (2002): Sensitivity of a coupled
is formed, in situ near-surface temperature measurement single-column model in the tropics to treatment of the in-
is not easy (Subsection 7.2). Careful measurements with- terfacial parameterization. J. Climate, 15, 18051831.
out instrument-induced turbulence and/or heating are re- Clayson, C. A. and J. A. Curry (1996): Determination of sur-
face turbulent fluxes for the Tropical Ocean-Global Atmos-
quired for studies of diurnal air-sea interaction.
phere Coupled Ocean-Atmosphere Response Experiment:
Again, the impact of the diurnal SST variability on
Comparison of satellite retrievals and in situ measurements.
air-sea interaction has not yet been investigated suffi- J. Geophys. Res., 101, 2851528528.
ciently. Even if the diurnal change of SST itself were to Clayson, C. A. and D. Weitlich (2005): Diurnal warming in the
have little effect on the atmosphere, it has been confirmed tropical Pacific and its interannual variability. Geophys. Res.
that the formation/decay process of the diurnal Lett., 32, L21604, doi:10.1029/2005GL023786.
thermocline near the surface cannot be neglected in or- Clayson, C. A. and D. Weitlich (2007): Variability of tropical
der to reproduce the SST variation on an intraseasonal or diurnal sea surface temperature. J. Climate, 20, 334352.
longer scale accurately in a model, which is essential to Cornillon, P. and L. Stramma (1985): The distribution of diur-
air-sea interaction, especially in the tropics. Further stud- nal sea surface warming events in the western Sargasso Sea.
ies using numerical models are necessary, and an observ- J. Geophys. Res., 90, 1181111815.
Cronin, M. F. and W. S. Kessler (2002): Seasonal and
ing system that can detect the diurnal cycle in SST all
interannual modulation of mixed layer variability at 0,
over the world is also indispensable for the corroboration
110W. Deep-Sea Res. I, 49, 117.
of the model simulation. Curry, J. A., A. Bentamy, M. A. Bourassa, D. Bourras, E. F.

740 Y. Kawai and A. Wada


Bradley, M. Brunke, S. Castro, S. H. Chou, C. A. Clayson, Wentz (2003): Diurnal signals in satellite sea surface tem-
W. J. Emery, L. Eymard, C. W. Fairall, M. Kubota, B. Lin, perature measurements. Geophys. Res. Lett., 30(3), 1140,
W. Perrie, R. A. Reeder, I. A. Renfrew, W. B. Rossow, J. doi:10.1029/2002GL016291.
Schulz, S. R. Smith, P. J. Webster, G. A. Wick and X. Zeng Gentemann, C. L., F. J. Wentz, C. A. Mears and D. K. Smith
(2004): Seaflux. Bull. Am. Meteorol. Soc., 85, 409423. (2004): In situ validation of Tropical Rainfall Measuring
Dai, A. and K. E. Trenberth (2004): The diurnal cycle and its Mission microwave sea surface temperature. J. Geophys.
depiction in the Community Climate System Model. J. Cli- Res., 109, C04021, doi:10.1029/2003JC002092.
mate, 17, 930951. Godfrey, J. S., R. A. Houze, Jr., R. H. Johnson, R. Lukas, J.-L.
Danabasoglu, G., W. G. Large, J. J. Tribbia, P. R. Gent, B. P. Redelsperger, A. Sumi and R. Weller (1998): Coupled
Briegleb and J. C. McWilliams (2006): Diurnal coupling in Ocean-Atmosphere Response Experiment (COARE): An
the tropical oceans of CCSM3. J. Climate, 19, 23472365. interim report. J. Geophys. Res., 103, 1439514450.
Deschamps, P. Y. and R. Frouin (1984): Large diurnal heating Gray, W. M. and R. W. Jacobson, Jr. (1977): Diurnal variation
of the sea surface observed by the HCMR experiment. J. of deep cumulus convestion. Mon. Wea. Rev., 105. 1171
Phys. Oceanogr., 14, 177184. 1188.
Deser, C. and C. A. Smith (1998): Diurnal and semidiurnal vari- Hepplewhite, C. L. (1989): Remote observation of the sea sur-
ations of the surface wind field over the tropical Pacific face and atmosphere. The oceanic skin effect. Int. J. Re-
Ocean. J. Climate, 11, 17301748. mote Sens., 10, 801810.
Dong, S., S. T. Gille, J. Sprintall and C. Gentemann (2006): Janowiak, J. E., P. A. Arkin and M. Morrissey (1994): An ex-
Validation of the Advanced Microwave Scanning Radiom- amination of the diurnal cycle in oceanic tropical rainfall
eter for the Earth Observing System (AMSR-E) sea surface using satellite and in situ data. Mon. Wea. Rev., 122, 2296
temperature in the Southern Ocean. J. Geophys. Res., 111, 2311.
C04002, doi:10.1029/2005JC002934. Jerlov, N. G. (1968): Optical Oceanography, Elsevier Ocea-
Donlon, C. J. and the GHRSST-PP Science Team (2005): The nography Series Vol. 5. Elsevier, Amsterdam, 194 pp.
Recommended GHRSST-PP Data Processing Specification Johnson, R. H., T. M. Rickenbach, S. A. Rutledge, P. E.
GDS (version 1 revision 1.6). The GHRSST-PP International Ciesielski and W. H. Schubert (1999): Trimodal character-
Project Office, Exeter, U.K., 245 pp. (available at: http:// istics of tropical convection. J. Climate, 12, 23972418.
ghrsst-pp.jrc.it/documents/GDS-v1.6.zip). Kahru, M., J.-M. Leppnen and O. Rud (1993): Cyanobacterial
Donlon, C. J., S. J. Keogh, D. J. Baldwin, I. S. Robinson, I. blooms cause heating of the sea surface. Mar. Ecol. Prog.
Ridley, T. Sheasby, I. J. Barton, E. F. Bradley, T. J. Nightin- Ser., 101, 17.
gale and W. J. Emery (1998): Solid-state radiometer meas- Kantha, L. H. and C. A. Clayson (1994): An improved mixed
urements of sea surface skin temperature. J. Atmos. Oce- layer model for geophysical applications. J. Geophys. Res.,
anic Technol., 15, 775787. 99, 2523525266.
Donlon C. J., P. J. Minnett, C. Gentemann, T. J. Nightingale, I. Katsaros, K. B. (1980): The aqueous thermal boundary layer.
J. Barton, B. Ward and M. J. Murray (2002): Toward im- Bound.-Layer Meteorol., 18, 107127.
proved validation of satellite sea surface skin temperature Kawai, Y. and H. Kawamura (1997): Seasonal and diurnal vari-
measurements for climate research. J. Climate, 15, 353 ability of differences between satellite-derived and in situ
369. sea surface temperatures in the south of the Sea of Okhotsk.
Donlon, C. J., I. S. Robinson, K. S. Casey, J. Vazquez-Cuervo, J. Oceanogr., 53, 343354.
E. Armstrong, O. Arino, C. L. Gentemann, D. May, P. Kawai, Y. and H. Kawamura (2000): Study on a platform effect
LeBorgne, J. Pioll, I. J. Barton, H. Beggs, D. J. S. Poulter, in the in situ sea surface temperature observations under
C. J. Merchant, A. Bingham, S. Heinz, A. Harris, G. A. weak wind and clear sky conditions using numerical mod-
Wick., W. J. Emery, P. J. Minnett, R. Evans, D. T. Llewellyn- els. J. Atmos. Oceanic Technol., 17, 185196.
Jones, C. T. Mutlow, R. W. Reynolds, H. Kawamura and N. Kawai, Y. and H. Kawamura (2002): Evaluation of the diurnal
A. Rayner (2007): The Global Ocean Data Assimilation warming of sea surface temperature using satellite-derived
Experiment High-resolution Sea Surface Temperature Pilot marine meteorological data. J. Oceanogr., 58, 805814.
Project. Bull. Am. Meteorol. Soc. (in press). Kawai, Y. and H. Kawamura (2003): Validation of daily ampli-
Emery, W. J., D. J. Baldwin, P. Schlessel and R. W. Reynolds tude of sea surface temperature evaluated with a paramet-
(2001): Accuracy of in situ sea surface temperatures used ric model using satellite data. J. Oceanogr., 59, 637644.
to calibrate infrared satellite measurements. J. Geophys. Kawai, Y. and H. Kawamura (2005): Spatial and temporal vari-
Res., 106, 23872405. ations of model-derived diurnal amplitude of sea surface
Fairall, C. W., E. F. Bradley, J. S. Godfrey, G. A. Wick, J. B. temperature in the western Pacific Ocean. J. Geophys. Res.,
Edson and G. S. Young (1996): Cool-skin and warm-layer 110, C08012, doi:10.1029/2004JC002652.
effects on sea surface temperature. J. Geophys. Res., 101, Kawai, Y., K. Otsuka and H. Kawamura (2006a): Study on di-
12951308. urnal sea surface warming and a local atmospheric circula-
Flament, P., J. Firing, M. Sawyer and C. Trefois (1994): Am- tion over Mutsu Bay. J. Meteorol. Soc. Japan, 84, 725744.
plitude and horizontal structure of a large diurnal sea sur- Kawai, Y., H. Kawamura, S. Tanba, K. Ando, K. Yoneyama and
face warming event during the Coastal Ocean Dynamics N. Nagahama (2006b): Validity of sea surface temperature
Experiment. J. Phys. Oceanogr., 24, 124139. observed with the TRITON buoy under diurnal heating con-
Gentemann, C. L., C. J. Donlon, A. Stuart-Menteth and F. J. ditions. J. Oceanogr., 62, 825838.

Diurnal SST Variation and Its Impact 741


Kearns, E. J., J. A. Hanafin, R. H. Evans, P. J. Minnett and O. Jones (1994): Sea surface temperature measurements by the
B. Brown (2000): An independent assessment of Pathfinder along-track scanning radiometer on the ERS 1 satellite:
AVHRR sea surface temperature accuracy using the Ma- Early results. J. Geophys. Res., 99, 2257522588.
rine Atmosphere Emitted Radiance Interferometer Nicls, R., V. Caselles, C. Coll, E. Valor and E. Rubio (2004):
(MAERI). Bull. Am. Meteorol. Soc., 81, 15251536. Autonomous measurements of sea surface temperature us-
Kilpatrick, K. A., G. P. Podest and R. Evans (2001): Over- ing in situ thermal infrared data. J. Atmos. Oceanic Technol.,
view of the NOAA/NASA advanced very high resolution 21, 683692.
radiometer Pathfinder algorithm for sea surface tempera- Noh, Y., H. S. Min and S. Raasch (2004): Large eddy simula-
ture and associated matchup database. J. Geophys. Res., 106, tion of the ocean mixed layer: The effects of wave breaking
91799197. and Langmuir circulation. J. Phys. Oceanogr., 34, 720735.
Kobayashi, T., K. Izawa, A. Inoue, N. Shikama, K. Ando, Y. Notarstefano, G., E. Mauri and P.-M. Poulain (2006): Near-sur-
Takatsuki, H. Nakajima, E. Oka, S. Hosoda, M. Miyazaki, face thermal structure and surface diurnal warming in the
N. Iwasaka, T. Suga, K. Mizuno, K. Takeuchi, all members Adriatic Sea using satellite and drifter data. Remote Sens.
of JAMSTEC/FORSGC Argo Team (2004): Solution of sa- Environ., 101, 194211.
linity-offset problem occurred in PROVOR type profiling Ohlmann, J. C. and D. A. Siegel (2000): Ocean radiant heating.
float. Report of Japan Marine Science and Technology Part II: Parameterizing solar radiation transmission through
Center, 49, 107120 (in Japanese with English abstract, the upper ocean. J. Phys. Oceanogr., 30, 18491865.
available at: http://www.jamstec.go.jp/pdf/shiken_pdf/49b/ Parsons, D. B., K. Yoneyama and J.-L. Redelsperger (2000):
11.pdf). The evolution of the tropical western Pacific atmosphere-
Koizumi, M. (1956): Researches on variations of oceanographic ocean system following the arrival of a dry intrusion. Q. J.
conditions in the region of the ocean weather station Ex- R. Meteorol. Soc., 126, 517548.
tra in the north Pacific Ocean (IV)On the diurnal varia- Paulson, C. A. and J. J. Simpson (1977): Irradiance measure-
tions in air and sea-surface temperatures. Pap. Meteor. ments in the upper ocean. J. Phys. Oceanogr., 7, 17221735.
Geophys., 7, 145154. Price, J. F., C. N. K. Moore and J. C. Van Leer (1978): Obser-
Kondo, J., Y. Sasano and T. Ishii (1979): On wind-driven cur- vation and simulation of storm-induced mixed-layer deep-
rent and temperature profiles with diurnal period in the oce- ening. J. Phys. Oceanogr., 8, 582599.
anic planetary boundary layer. J. Phys. Oceanogr., 9, 360 Price, J. F., R. A. Weller and R. Pinkel (1986): Diurnal cycling:
372. Observations and models of upper ocean response to diur-
Kubota, M., A. Kano, H. Muramatsu and H. Tomita (2003): nal heating, cooling, and wind mixing. J. Geophys. Res.,
Intercomparison of various surface latent heat flux fields. 91, 84118427.
J. Climate, 16, 670678. Price, J. F., R. A. Weller, C. M. Bowers and M. G. Briscoe
Large, W. G., J. C. McWilliams and S. C. Doney (1994): Ocean (1987): Diurnal response of sea surface temperature ob-
vertical mixing: A review and a model with a nonlocal served at the long-term upper ocean study (34N, 70W) in
boundary layer parameterization. Rev. Geophys., 32, 363 the Sargasso Sea. J. Geophys. Res., 92, 1448014490.
403. Randall, D. A., Harshvardhan and D. A. Dazlich (1991): Diur-
Lee, T. and W. T. Liu (2005): Effects of high-frequency wind nal variability of the hydrological cycle in a general circu-
sampling on simulated mixed layer depth and upper ocean lation model. J. Atmos. Sci., 48, 4062.
temperature. J. Geophys. Res., 110, C05002, doi:10.1029/ Rayner, N. A., D. E. Parker, E. B. Horton, C. K. Folland, L. V.
2004JC002746. Alexander, D. P. Rowell, E. C. Kent and A. Kaplan (2003):
Li, W., R. Yu, H. Liu and Y. Yu (2001): Impacts of diurnal cy- Global analyses of sea surface temperature, sea ice, and
cle of SST on the interseasonal variation of surface heat night marine air temperature since the late nineteenth cen-
flux over the western Pacific warm pool. Adv. Atmos. Sci., tury. J. Geophys. Res., 108(D14), 4407, doi:10.1029/
18, 793806. 2002JD002670.
Mapes, B. E., T. T. Warner and M. Xu (2003): Diurnal pattern Reynolds, R. W. and T. M. Smith (1994): Improved global sea
of rainfall in northwestern South America. Part III: Diurnal surface temperature analyses using optimum interpolation.
gravity waves and nocturnal convection offshore. Mon. Wea. J. Climate, 7, 929948.
Rev., 131, 830844. Reynolds, R. W., N. A. Rayner, T. M. Smith, D. C. Stokes and
McCreary, J. P., K. E. Kohler, R. R. Hood, S. Smith, J. Kindle, W. Wang (2002): An improved in situ and satellite SST
A. S. Fischer and R. A. Weller (2001): Influence of diurnal analysis for climate. J. Climate, 15, 16091625.
cycle and intraseasonal forcing on mixed layer and biologi- Riser, S. and W. Wijffels (2005): Report from the 1st Argo Tech-
cal variability in the central Arabian Sea. J. Geophys. Res., nical Workshop. University of Washington, Seattle, 28 pp.
106, 71397155. Robertson, J. E. and A. J. Watson (1992): Thermal skin effect
McNeil, C. L. and L. Merlivat (1996): The warm oceanic sur- of the surface ocean and its implication for CO2 uptake.
face layer: Implications for CO 2 fluxes and surface gas Nature, 358, 738740.
measurements. Geophys. Res. Lett., 23, 35753578. Robinson, I. S., N. C. Wells and H. Charnock (1984): The sea
Mellor, G. L. and T. Yamada (1982): Development of a turbu- surface thermal boundary layer and its relevance to the
lence closure model for geophysical fluid problems. Rev. measurement of sea surface temperature by airborne and
Geophys. Space Phys., 20, 851875. spaceborne radiometer. Int. J. Remote Sens., 5, 1945.
Mutlow, C. T., A. M. Zvody, I. J. Barton and D. T. Llewellyn- Roll, H. U. (1965): Physics of the Marine Atmosphere, Inter-

742 Y. Kawai and A. Wada


national Geophysics Series Vol. 7. Academic Press, New GasEx-98. p. 181185. In Gas Transfer at Water Surfaces,
York, 426 pp. Geophys. Monogr., No. 127, ed. by M. A. Donelan, W. M.
Sakaida, F., J. Kudoh and H. Kawamura (2000): A-HIGHERS Drennan, E. S. Saltzmann and R. Wanninkhov, Amer.
The system to produce the high spatial resolution sea sur- Geophys. Union, Washington, D.C.
face temperature maps of the western North Pacific using Stommel, H. and A. H. Woodcock (1951): Diurnal heating of
the AVHRR/NOAA. J. Oceanogr., 56, 707716. the surface of the Gulf of Mexico in the spring of 1942.
Sarmiento, J. L. and E. T. Sundquist (1992): Revised budget Trans. Am. Geophys. Un., 32, 565571.
for the oceanic uptake of anthropogenic carbon dioxide. Stommel, H., K. Saunders, W. Simmons and J. Cooper (1969):
Nature, 356, 589593. Observations of the diurnal thermocline. Deep-Sea Res., 16,
Saunders, P. M. (1967): The temperature at the ocean-air inter- 269284.
face. J. Atmos. Sci., 24, 269273. Stramma, L., P. Cornillon, R. A. Weller, J. F. Price and M. G.
Saur, T. F. T. (1963): A study of the equality of sea water tem- Briscoe (1986): Large diurnal sea surface temperature vari-
peratures reported in logs of ships weather observations. ability: Satellite and in situ measurements. J. Phys.
J. Appl. Meteorol., 2, 417425. Oceanogr., 16, 827837.
Schiller, A. and J. S. Godfrey (2003): Indian Ocean intraseasonal Stuart-Menteth, A. C., I. S. Robinson and P. G. Challenor
variability in an Ocean General Circulation Model. J. Cli- (2003): A global study of diurnal warming using satellite-
mate, 16, 2139. derived sea surface temperature. J. Geophys. Res., 108(C5),
Schiller, A. and J. S. Godfrey (2005): A diagnostic model of 3155, doi:10.1029/2002JC001534.
the diurnal cycle of sea surface temperature for use in cou- Subrahmanyam, B., K. H. Rao, N. S. Rao, V. S. N. Murty and
pled ocean-atmosphere models. J. Geophys. Res., 110, R. J. Sharp (2002): Influence of a tropical cyclone on chlo-
C11014, doi:10.1029/2005JC002975. rophyll-a concentration in the Arabian Sea. Geophys. Res.
Schlessel, P., H.-Y. Shin, W. J. Emery and H. Grassl (1987): Lett., 29(22), 2065, doi:10.1029/2002GL015892.
Comparison of satellite-derived sea surface temperatures Sui, C.-H., K.-M. Lau, Y. N. Takayabu and D. A. Short (1997a):
with in situ skin measurements. J. Geophys. Res., 92, 2859 Diurnal variations in tropical oceanic cumulus cloud con-
2874. vection during TOGA COARE. J. Atmos. Sci., 54, 639655.
Schlessel, P., W. J. Emery, H. Grassl and T. Mammen (1990): Sui, C.-H., X. Li, K.-M. Lau and D. Adamec (1997b): Multiscale
On the bulk-skin temperature difference and its impact on air-sea interactions during TOGA COARE. Mon. Wea. Rev.,
satellite remote sensing of sea surface temperature. J. 125, 448462.
Geophys. Res., 95, 1334113355. Sverdrup, H. U., M. W. Johnson and R. H. Fleming (1942): The
Shinoda, T. (2005): Impact of the diurnal cycle of solar radia- Oceans: Their Physics, Chemistry and General Biology.
tion on intraseasonal SST variability in the western equato- Prentice-Hall, Englewood Cliffs, New York, 1087 pp.
rial Pacific. J. Climate, 18, 26282636. Tachibana, Y., H. Ogawa, K. Iwamoto, K. Takeuchi and M.
Shinoda, T. and H. H. Hendon (1998): Mixed layer modeling Wakatsuchi (2004): An observational study on the marine
of intraseasonal variability in the tropical western Pacific fog associated with the Okhotsk high. Gekkan Kaiyo, 36,
and Indian Oceans. J. Climate, 11, 26682685. 305308 (in Japanese).
Siegel, D. A., J. C. Ohlmann, L. Washburn, R. R. Bidigare, C. Tanahashi, S., H. Kawamura, T. Takahashi and H. Yusa (2003):
T. Nosse, E. Fields and Y. Zhou (1995): Solar radiation, Diurnal variations of sea surface temperature over the wide-
phytoplankton pigments and the radiant heating of the equa- ranging ocean using VISSR on board GMS. J. Geophys.
torial Pacific warm pool. J. Geophys. Res., 100, 48854891. Res., 108(C7), 3216, doi:10.1029/2002JC001313.
Siswanto, E., J. Ishizaka, K. Yokouchi, K. Tanaka and C. K. Van Scoy, K. A., K. P. Morris, J. E. Robertson and A. J. Watson
Tan (2007): Estimation of interannual and interdecadal vari- (1995): Thermal skin effect and the air-sea flux of carbon
ations of typhoon-induced primary production: A case study dioxide: A seasonal high-resolution estimate. Global
for the outer shelf of the East China Sea. Geophys. Res. Biogeochem. Cycles, 9, 253262.
Lett., 34, L03604, doi:10.1029/2006GL028368. Ward, B. (2006): Near-surface ocean temperature. J. Geophys.
Slingo, J. M., P. Inness, R. Neale, S. Woolnough and G.-Y. Yang Res., 111, C02005, doi:10.1029/2004JC002689.
(2003): Scale interactions on diurnal to seasonal timescales Ward, B. and M. A. Donelan (2006): Thermometric measure-
and their relevance to model systematic errors. Ann. ments of the molecular sublayer at the air-water interface.
Geophys., 46, 139155. Geophys. Res. Lett., 33, L07605, doi:10.1029/
Soloviev, A. and R. Lukas (1997): Observation of large diurnal 2005GL024769.
warming events in the near-surface layer of the western Ward, B. and P. J. Minnett (2001): An autonomous profiler for
equatorial Pacific warm pool. Deep-Sea Res. I, 44, 1055 near surface temperature measurements. p. 167172. In Gas
1076. Transfer at Water Surfaces, Geophys. Monogr., No. 127,
Soloviev, A. and R. Lukas (2006): The near-surface layer of ed. by M. A. Donelan, W. M. Drennan, E. S. Saltzmann and
the ocean: Structure, dynamics and application. Atmospheric R. Wanninkhov, Amer. Geophys. Union, Washington, D.C.
and Oceanographic Sciences Library Vol. 31. Springer, Ward, B., R. Wanninkhof, W. R. McGillis, A. T. Jessup, M. D.
Dordrecht, 572 pp. DeGrandpre, J. E. Hare and J. B. Edson (2004a): Biases in
Soloviev, A., J. Edson, W. McGillis, P. Schlessel and R. the air-sea flux of CO 2 resulting from ocean surface
Wanninkof (2001): Fine thermohaline structure and gas- temeperature gradients. J. Geophys. Res., 109, C08S08,
exchange in the near-surface layer of the ocean during doi:10.1029/2003JC001800.

Diurnal SST Variation and Its Impact 743


Ward, B., R. Wanninkhof, P. J. Minnett and M. J. Head (2004b): Worley, S. J., S. D. Woodruff, R. W. Reynolds, S. J. Lubker and
SkinDeEP: A profiling instrument for upper-decameter sea N. Lott (2005): ICOADS release 2.1 data and products. Int.
surface measurements. J. Atmos. Oceanic Technol., 21, 207 J. Climatol., 25, 823842.
222. Wu, X., W. P. Menzel and G. S. Wade (1999): Estimation of sea
Webster, P. J., C. A. Clayson and J. A. Curry (1996): Clouds, surface temperatures using GOES-8/9 radiance measure-
radiation, and the diurnal cycle of sea surface temperature ments. Bull. Am. Meteorol. Soc., 80, 11271138.
in the tropical western Pacific. J. Climate, 9, 17121730. Yang, G.-Y. and J. Slingo (2001): The diurnal cycle in the trop-
Weller, R. A. and S. P. Anderson (1996): Surface meteorology ics. Mon. Wea. Rev., 129, 784801.
and air-sea fluxes in the western equatorial Pacific warm Yokoyama, R., S. Tanba and T. Souma (1995): Sea surface ef-
pool during the TOGA Coupled Ocean-Atmosphere Re- fects on the sea surface temperature estimation by remote
sponse Experiment. J. Climate, 9, 19591990. sensing. Int. J. Remote Sens., 16, 227238.
Wick, G. A., J. J. Bates and D. J. Scott (2002): Satellite and Yu, L., R. A. Weller and B. Sun (2004): Mean and variability of
skin-layer effects on the accuracy of sea surface tempera- the WHOI daily latent and sensible heat fluxes at in situ
ture measurements from the GOES satellites. J. Atmos. flux measurement sites in the Atlantic Ocean. J. Climate,
Oceanic Technol., 19, 18341848. 17, 20962118.
Wick, G. A., J. C. Ohlmann, C. W. Fairall and A. T. Jessup Zhang, C. (2005): Madden-Julian Oscillation. Rev. Geophys.,
(2005): Improved oceanic cool-skin corrections using a re- 43, RG2003, doi10.1029/2004RG000158.
fined solar penetration model. J. Phys. Oceanogr., 35, 1986 Zeng, X. and A. Beljaars (2005): A prognostic scheme of sea
1996. surface skin temperature for modeling and data assimila-
Wong, C. S., Y.-H. Chan and J. S. Page (1995): Geographical, tion. Geophys. Res. Lett., 32, L14605, doi:10.1029/
seasonal and interannual variations of air-sea CO2 exchange 2005GL023030.
in the subtropical Pacific surface waters during 19831988. Zeng, X. and R. E. Dickinson (1998): Impact of diurnally-vary-
(II) Air-sea CO2 fluxes with skin-temperature adjustments. ing skin temperature on surface fluxes over the tropical
Tellus, 47B, 431446. Pacific. Geophys. Res. Lett., 25, 14111414.
Woolnough, S. J., F. Vitart and M. A. Balmaseda (2007): The Zeng, X., M. Zhao, R. E. Dickinson and Y. He (1999): A
role of the ocean in the Madden-Julian Oscillation: Impli- multiyear hourly sea surface skin temperature data set de-
cations for MJO prediction. Q. J. R. Meteorol. Soc., 133, rived from the TOGA TAO bulk temperature and wind speed
117128. over the tropical Pacific. J. Geophys. Res., 104, 15251536.

744 Y. Kawai and A. Wada

You might also like