grl51416-sup-0001

advertisement
Modeled Ross Sea Changes
1
SUPPLEMENTAL MATERIAL
2
1. Model setup and simulations
3
The Ross Sea model [Dinniman et al., 2007, 2011] now includes a dynamic sea ice model
4
[Budgell, 2005] that prognostically calculates sea-ice concentration and thickness [Stern et al.,
5
2013] and accurately depicts circulation and shelf water formation [Dinniman et al., 2007, 2011].
6
Open-ocean momentum, as well as heat and fresh water (imposed as a salt flux) fluxes, are
7
calculated based on the COARE 3.0 bulk flux algorithm [Fairall et al., 2003] with no relaxation
8
of surface temperature or salinity. The Ross Sea model also simulates the mechanical and
9
thermodynamic interactions between the floating Ross Ice Shelf and the water underneath
10
[Holland and Jenkins, 1999; Dinniman et al., 2011]. Inputs along the open boundaries of the
11
regional model [Dinniman et al., 2007] have been updated to use sea-ice concentrations derived
12
from Special Sensor Microwave/Imager (SSM/I). Ocean tides are not included. However, this
13
model has now been run with tides [Padman et al., in prep; Mack et al., in prep] and the addition
14
of tides does not have a significant impact on the aspects of the circulation discussed here.
15
High resolution (30-km grid spacing) winds and atmospheric temperatures were obtained
16
from the Antarctic Mesoscale Prediction System [AMPS; Powers et al., 2003; Bromwich et al.,
17
2005]. Atmospheric forecasts for September 2003 to September 2005 were used to force the
18
Ross Sea circulation model. The forcing fields resulted in accurate simulation of the maximum
19
summer extent of the Ross Sea polynya (Fig. S1). Similar results were obtained with the ERA-
20
Interim winds (0.75° horizontal spacing) for this and other time periods (see results in Stern et al.
21
[2013]). Simulations that used the CMIP3 MPI-ECHAM5 winds (1.875° horizontal spacing)
22
produced less accurate representations of the summer polynya extent, suggesting that resolution
23
in the wind forcing is a critical aspect of simulating sea ice distribution in the Ross Sea.
1
Modeled Ross Sea Changes
24
Daily winds and atmospheric temperature fields for 9/1996-9/2000 were obtained from the
25
MPI ECHAM5 climate model [Jungclaus et al., 2006] 20th-century experiment. In an evaluation
26
of the IPCC AR4 climate model simulations, Connolley and Bracegirdle [2007] concluded that
27
MPI ECHAM5 was one of the two best performers in the Antarctic (and best globally), and as a
28
result we used this atmospheric model. MPI ECHAM5 was also one of the CMIP3 models that
29
included 21st century stratospheric ozone recovery [Miller et al., 2006]. Simulations using a
30
global sea ice-ocean model forced with future emissions scenarios from ECHAM5 or the Hadley
31
Centre Climate Model (the other top performer in the Antarctic) showed little difference between
32
the two in sea ice extent, although there were significant differences in other aspects of the ocean
33
simulation [Timmermann and Hellmer, 2013]. Comparison of these atmospheric temperature
34
fields with the ECMWF-Interim (ERA-Int) Reanalysis temperature data showed that the
35
ECHAM5 temperatures were colder (summer climatology differences of up to 8 °C), especially
36
near the coast and Ross Ice Shelf front. Simulations using the 20th-century temperatures did not
37
produce significant summer expansion of the Ross Sea Polynya (not shown). Therefore, a
38
monthly climatology constructed from the ERA-Int air temperatures was used as input for the
39
20th-century simulations.
40
Daily winds for 9/2046-9/2050 and 9/2096-9/2100 were taken from the CMIP3 A1B
41
emissions scenario of the ECHAM5 model. Monthly climatologies of the difference between the
42
ECHAM5 A1B and the 20th-century air temperatures were computed for both future scenarios,
43
and added to the ERA-Int climatology to create forcing fields for the ocean model. The adjusted
44
air temperatures were 1.66 ± 2.58°C warmer for 2046-2050 (Year 2050) and 2.50 ± 2.32°C
45
warmer for 2096-2100 (Year 2100) relative to the 20th-century air temperatures.
2
Modeled Ross Sea Changes
46
The observed freshening of the Ross Sea [salinity changes of -0.03 decade-1 near Ross Island,
47
-0.04 decade-1 in MCDW near the Ross Ice Shelf, and -0.08 decade-1 above 200 m along the
48
Ross Ice Shelf and near the slope front have been observed; Jacobs and Giulivi, 2010] was
49
simulated using a boundary condition that uniformly reduced salinity at the boundaries by 0.12.
50
While this value of freshening may not be realistic on the model’s northern or western
51
boundaries, as is it believed that most of the freshening is driven from upstream in the
52
Amundsen-Bellingshausen Seas [Jacobs and Giulivi,2010], the largest transport into the model
53
domain from outside is the shelf break current at the eastern boundary, and thus the only part of
54
the boundary freshening that has a significant effect on the simulation is the setup on the eastern
55
boundary. Simulations that used the 2050 or 2100 winds and adjusted air temperature, but
56
without freshening were done to isolate the effects of wind and air temperature on oceanographic
57
properties (Table S1).
58
2. Simulated sea-ice area and influence on productivity
59
Observed and simulated sea-ice concentrations over the Ross Sea continental shelf are near
60
100% in winter, but decrease substantially with the seasonal development of the Ross Sea
61
polynya (Fig. 2). Under the 2050 scenario, winter sea-ice concentrations are similar to present
62
conditions, but the minimum summer sea-ice cover decreases to a mean ice area that is 44% of
63
present conditions. In 2100 the summer ice cover is only 22% of that at present.
64
The change in the mass flux between the ocean and the sea ice over the continental shelf
65
(Fig. S2) suggests a significant increase in melting (negative mass flux) for the 2050 and 2100
66
scenarios, indicating that much of the difference in the summer ice-free extent is due to in situ
67
melting rather than enhanced advection of sea ice out of the Ross Sea. Changes in the boundary
68
salinity (e.g., 2050SBC vs. 2050) made little difference in the local melting/freezing patterns.
3
Modeled Ross Sea Changes
69
The timing of the greatest in situ melting is earlier in 2050 than present (Fig. S2) and becomes
70
even earlier in 2100. This represents a shift in the timing of the polynya formation, with the
71
mean day of retreat below 15% concentration (calculated in the same manner as in Stammerjohn
72
et al. [2008]) over the continental shelf in 2050 occurring 5 days earlier, but the mean day of
73
advance occurring at the same time . Some locations have a shorter ice-free period due to a
74
slight shift in the location of the summer polynya (Fig. S3a), suggesting the importance of spatial
75
variability in the changes. By 2100, the duration of the ice-free season is even longer (Fig. S3b)
76
with the mean day of retreat now occurring 11 days earlier and the advance occurring 16 days
77
later.
78
The simulated changes in the ice-free period were used to estimate summer mean primary
79
productivity, which was added to annual mean primary productivity estimates to determine
80
potential effects on the biological production of the Ross Sea (Table 1). Diatoms frequently
81
dominate summer production in the Ross Sea [Smith et al., 2011], and this contribution is also
82
estimated (Table 1). The Ross Sea productivity estimates [Smith et al., 2014] are based on
83
nitrogen and silicon nutrient deficits from a limited number of samples taken in different years.
84
Total productivity is comparable to satellite-derived estimates [Arrigo et al., 2008].
85
3. Location of differences in surface layer CDW dye
86
A substantial volume of MCDW is advected onto the continental shelf along the troughs
87
(Fig.3), where it is mixed into the upper water column (Fig. S4a). In the 2050 and 2100
88
simulations, the amount of MCDW that enters the upper 50 m of the water column (averaged
89
over the entire continental shelf) decreased by 24 and 8%, respectively (Fig. S4b,c). More
90
CDW/MCDW dye gets mixed into the surface layers in the western Ross Sea, which presently is
4
Modeled Ross Sea Changes
91
the site of the greatest vertical mixing and thus the location of the greatest changes in the surface
92
CDW inputs (Fig. S4).
93
94
5
Modeled Ross Sea Changes
95
References
96
Arrigo, K.R., G.L. van Dijken, and S. Bushinsky (2008), Primary production in the Southern
97
98
99
100
101
102
Ocean,1997-2006. J. Geophys. Res.,113, C08004, doi:10.1029/2007JC004551.
Bromwich, D.H., A.J. Monaghan, K.W. Manning, and J.G. Powers (2005), Real-time forecasting
for the Antarctic: An evaluation of the Antarctic Mesoscale Prediction System (AMPS).
Mon. Wea. Rev., 133, 579-603.
Budgell, W.P. (2005), Numerical simulation of ice-ocean variability in the Barents Sea region
towards dynamical downscaling. Ocean Dyn., 55, 370-387.
103
Dinniman, M.S., J.M. Klinck, and W.O. Smith, Jr. (2007), The influence of sea ice cover and
104
icebergs on circulation and water mass formation in a numerical circulation model of the
105
Ross Sea, Antarctica. J. Geophys. Res., 112, C11013, doi:10.1029/2006JC004036.
106
Dinniman, M.S., J.M. Klinck, and W.O. Smith, Jr. (2011), A model study of Circumpolar Deep
107
Water on the West Antarctic Peninsula and Ross Sea continental shelves. Deep-Sea Res. II,
108
58, 1508-1523.
109
Fairall, C.W., E.F. Bradley, J.E. Hare, A.A. Grachev, and J.B. Edson (2003), Bulk
110
parameterization of air-sea fluxes: Updates and verification for the COARE algorithm. J.
111
Climate, 16, 571-591.
112
113
114
115
116
117
Holland, D.M., and A. Jenkins (1999), Modeling thermodynamic ice-ocean interactions at the
base of an ice shelf. J. Phys. Oceanogr., 29, 1787-1800.
Jacobs, S.S., and C.F. Giulivi (2010), Large multidecadal salinity trends near the PacificAntarctic continental margin. J. Climate, 23, 4508-4524, doi:10.1175/2020JCLI3284.1.
Jungclaus, J.H. et al. (2006), Ocean circulation and tropical variability in the coupled model
ECHAM5/MPI-OM. J. Climate, 19, 3952-3972.
6
Modeled Ross Sea Changes
118
Miller, R.L., G.A. Schmidt, and D.T. Shindell (2006), Forced annular variations in the 20th
119
century Intergovernmental Panel on Climate Change Fourth Assessment Report models. J.
120
Geophys. Res., 111, D18101, doi:10.1029/2005JD006323.
121
Powers, J.G., A.J. Monaghan, A.M. Cayette, D.H. Bromwich, Y.-H. Kuo, and K.W. Manning
122
(2003), Real-time mesoscale modeling over Antarctica: The Antarctic Mesoscale Prediction
123
System (AMPS). Bull. Amer. Meteor. Soc., 84, 1533-1545.
124
Smith, W.O. Jr, D.G. Ainley, K.R. Arrigo, and M.S. Dinniman, M.S. (2014), The oceanography
125
and ecology of the Ross Sea. Ann. Rev. Mar. Sci., 6, 469-487, doi:10.1146/annurev-marine-
126
010213-135114 .
127
Smith, W.O. Jr., V. Asper, S. Tozzi, X. Liu, X. and S.E. Stammerjohn (2011), Continuous
128
fluorescence measurements in the Ross Sea, Antarctica: scales of variability. Prog.
129
Oceanogr., 88, 28–45.
130
Stammerjohn, S.E., D.G. Martinson, R.C. Smith, X. Yuan, and D. Rind (2008), Trends in
131
Antarctic annual sea ice retreat and advance and their relation to El Niño-Southern
132
Oscillation and Southern Annular Mode variability. J. Geophys. Res., 113, C03S90,
133
doi:10.1029/2007JC004269.
134
Stern, A.A., M.S. Dinniman, V. Zagorodnov, S.W. Tyler, and D.M. Holland (2013), Intrusion of
135
warm surface water beneath the McMurdo Ice Shelf, Antarctica. J. Geophys. Res., 118, 1-
136
13, doi:10.1002/2013JC008842.
7
Modeled Ross Sea Changes
137
Supplemental Figure Legends
138
S1. Simulated sea ice cover over the Ross Sea continental shelf obtained using 30-km resolution
139
AMPS winds and air temperatures for September 2003 to September 2005. The observed
140
(SSM/I) sea-ice cover for the same two-year period is shown for comparison.
141
S2. Averages of total mass flux (m3 water s-1), including frazil ice production, between the ocean
142
and sea ice on the Ross Sea continental shelf over the four years of each of the five
143
simulations.
144
S3. Simulated change in the number of days when sea ice concentrations are below 15%
145
obtained from the difference in a) the 2050 and the 20th century simulations and b) the 2100
146
and the 20th century simulations. Positive values correspond to a longer sea ice-free period.
147
S4. Concentration (in arbitrary units; a.u.) of CDW dye in the model surface layer at the end of
148
four-year simulations that used conditions for a) present, b) 2050, and c) 2100.
149
150
8
Download