0807.4180/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: 
3: \pdfoutput=1
4: 
5: %\usepackage{float}
6: \usepackage{epsfig}
7: %\usepackage{emulateapj5}
8: %\usepackage{onecolfloat5}
9: %\usepackage{apjfonts}
10: \usepackage{natbib}
11: \bibliographystyle{apj}
12: 
13: \def\plotone#1{\centering \leavevmode
14: \includegraphics[clip=, width=.85\columnwidth]{#1}}
15: \def\plottwo#1#2{\centering \leavevmode
16: \includegraphics[width=.45\columnwidth]{#1} \hfil
17: \includegraphics[width=.45\columnwidth]{#2}}
18: \def\plotthree#1#2#3{\centering \leavevmode
19: \includegraphics[width=.3\columnwidth]{#1} \hfil
20: \includegraphics[width=.3\columnwidth]{#2} \hfil
21: \includegraphics[width=.3\columnwidth]{#3}}
22: 
23: \newcommand{\cN}[1]{\mathcal{N}}
24: \newcommand{\pn}[1]{\mbox{$(#1)$}}
25: \newcommand{\spa}{\mbox{ }}
26: \def\gsim{\;\rlap{\lower 2.5pt
27:  \hbox{$\sim$}}\raise 1.5pt\hbox{$>$}\;}
28: \def\lsim{\;\rlap{\lower 2.5pt
29:    \hbox{$\sim$}}\raise 1.5pt\hbox{$<$}\;}
30: 
31: % set formatting properties
32: \setlength{\textwidth}{6.5in}
33: \setlength{\textheight}{9in}
34: \setlength{\hoffset}{0.in}
35: \setlength{\voffset}{-0.33in}
36: %\setlength{\voffset}{0.3in}
37: \parindent 0.0in
38: \parskip 0.2in
39: 
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41: %    % This is for the draft version watermark
42: %    \newcommand{\reviewtimetoday}[2]{\special{!userdict begin
43: %    /bop-hook{gsave 22 650 translate 60 rotate 0.8 setgray
44: %%    /bop-hook{gsave 20 710 translate 45 rotate 0.8 setgray
45: %    /Courier findfont 12 scalefont setfont 0 0   moveto (#1) show
46: %%    /Times-Roman findfont 12 scalefont setfont 0 0   moveto (#1) show
47: %    0 -12 moveto (#2) show grestore}def end}}
48: %    % You can turn on or off this option.
49: %    \reviewtimetoday{\today}{Draft Version}
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: 
52: 
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54: % THE DOCUMENT BEGINS HERE                      %
55: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
56: 
57: \begin{document}
58: 
59: 
60: %%% Begin front material
61: %\twocolumn[%%% Begin front material
62: 
63: \title{Habitable Climates: The Influence of Obliquity}
64: 
65: \author{David S. Spiegel\altaffilmark{1,2}, Kristen Menou\altaffilmark{1},
66: Caleb A. Scharf\altaffilmark{1,3}}
67: 
68: \affil{$^1$Department of Astronomy, Columbia University, 550 West 120th Street, New York, NY 10027}
69: \affil{$^2$Department of Astrophysical Sciences, Princeton University, Peyton Hall, Princeton, NJ 08544}
70: \affil{$^3$Columbia Astrobiology Center, Columbia University, 550 West 120th Street, New York, NY 10027}
71: 
72: \vspace{0.5\baselineskip}
73: 
74: \email{dave@astro.columbia.edu, kristen@astro.columbia.edu,
75:   caleb@astro.columbia.edu}
76: 
77: 
78: 
79: 
80: \begin{abstract}
81: Extrasolar terrestrial planets with the potential to host life might
82: have large obliquities or be subject to strong obliquity
83: variations. We revisit the habitability of oblique planets with an
84: energy balance climate model (EBM) allowing for dynamical transitions
85: to ice-covered snowball states as a result of ice-albedo
86: feedback. Despite the great simplicity of our EBM, it captures
87: reasonably well the seasonal cycle of global energetic fluxes at
88: Earth's surface.  It also performs satisfactorily against a
89: full-physics climate model of a highly oblique Earth--like planet, in
90: an unusual regime of circulation dominated by heat transport from the
91: poles to the equator.  Climates on oblique terrestrial planets can
92: violate global radiative balance through much of their seasonal cycle,
93: which limits the usefulness of simple radiative equilibrium
94: arguments. High obliquity planets have severe climates, with large
95: amplitude seasonal variations, but they are not necessarily more prone
96: to global snowball transitions than low obliquity planets. We find
97: that terrestrial planets with massive $\rm CO_2$ atmospheres,
98: typically expected in the outer regions of habitable zones, can also
99: be subject to such dynamical snowball transitions. Some of the
100: snowball climates investigated for $\rm CO_2$--rich atmospheres
101: experience partial atmospheric collapse. Since long-term $\rm CO_2$
102: atmospheric build-up acts as a climatic thermostat for habitable
103: planets, partial $\rm CO_2$ collapse could limit the habitability of
104: such planets. A terrestrial planet's habitability may thus depend
105: sensitively on its short-term climatic stability.
106: \end{abstract}
107: 
108: \keywords{astrobiology -- planetary systems --
109: radiative transfer}%]%%% End front material
110: 
111: \section{Introduction}
112: \label{obl_sec:intro}
113: 
114: The Earth's obliquity is remarkably stable: the angle between the
115: spin--axis and the normal to the orbital plane varies by no more than
116: a few degrees from its present value of $\sim 23.5\degr$.  This
117: stability is maintained by torque from the Moon
118: \citep{laskar_et_al1993,nerondesurgy+laskar1997}.  Even within our own
119: Solar System, though, the obliquity of other terrestrial planets has
120: varied significantly more; the analysis of \citet{laskar+robutel1993}
121: indicates that Mars' obliquity exhibits chaotic variations between
122: $\sim 0\degr$ and $\sim 60\degr$.
123: 
124: %\clearpage
125: 
126: How does climate depend on obliquity and its possible variations in
127: time?  How does the range of orbital radii around a star at which a
128: planet could support water-based life depend on the planet's
129: obliquity?  Has the stability of Earth's obliquity made it a more
130: climatically hospitable home?  The answers to these questions will be
131: important to evaluate the fraction of stars that have potentially
132: habitable planets.  There are now more than 300 extrasolar planets
133: known,\footnote{See http://exoplanet.eu/.} several of which are close
134: to the terrestrial regime with masses less than 10 times that of the
135: Earth (e.g., \citealt{beaulieu_et_al2006}, \citealt{udry_et_al2007},
136: \citealt{bennett_et_al2008}, \citealt{mayor_et_al2008}).  {\it COROT},
137: which has already launched, and {\it Kepler}, scheduled to launch in
138: less than one year, are dedicated space-based transit-detecting
139: observatories that will monitor a large number of stars to detect the
140: small decreases in stellar flux that occur when terrestrial planets
141: cross in front of their host stars
142: \citep{baglin2003,borucki_et_al2003,borucki_et_al2007}.  These
143: missions are expected to multiply by perhaps several hundredfold or
144: more the number of known terrestrial planets, depending on the
145: distribution of such planets around solar-type stars
146: (\citealt{borucki_et_al2007, borucki_et_al2003, basri_et_al2005};
147: although see revised predictions in \citealt{beatty+gaudi2008}).  NASA
148: and ESA have plans for ambitious future missions to obtain spectra of
149: nearby Earth-like planets in the hope that they would reveal the first
150: unambiguous signatures of life on a remote world: NASA's {\it
151: Terrestrial Planet Finder} and ESA's {\it Darwin}
152: \citep{leger+herbst2007}.  The design of such observatories, and the
153: urgency with which they will be built and deployed, will depend on the
154: habitability potential of terrestrial planets that will be found in
155: the next 5-10 years.
156: 
157: Over the last 50 years, various authors have addressed how to predict
158: the way in which terrestrial planet habitability depends on
159: star-planet distance (see \citealt{kasting+catling2003} for a recent
160: review).  Several of the important initial calculations predated the
161: first discoveries of extrasolar planets, including \citet{dole1964},
162: \citet{hart1979}, and the seminal work of \citet{kasting_et_al1993}.
163: \citet{selsis_et_al2007}, \citet{vonbloh_et_al2008}, and
164: \citet{barnes_et_al2008} have reconsidered habitability in light of
165: recent exoplanetary detections.  \citet{williams_et_al1996},
166: \citet[hereafter WK97]{williams+kasting1997} and
167: \citet{williams+pollard2003} have tackled precisely the questions
168: relating to obliquity posed above, and have concluded that variations
169: in obliquity do not necessarily render a planet non-habitable (see
170: also \citealt{hunt1982}, \citealt{williams1988b,williams1988c},
171: \citealt{oglesby+ogg1998}, \citealt{chandler+sohl2000} and
172: \citealt{jenkins2000,jenkins2001,jenkins2003} in the context of
173: Earth's paleoclimate studies).
174: 
175: Here we seek to generalize these analyses to model planets that are
176: less close analogs to Earth than have been considered previously.  In
177: \citet[hereafter SMS08]{spiegel_et_al2008}, we examined how regionally
178: and temporally habitable climates are affected by variations in the
179: efficiency of latitudinal heat transport on a planet, and by
180: variations in the ocean fraction. Importantly, we found that otherwise
181: habitable Earth-like terrestrial planets can be subject to dynamical
182: climate transitions into globally-frozen snowball states. Since it is
183: not trivial to escape a snowball state (e.g.,
184: \citealt{pierrehumbert2005}) and such globally-frozen climates may
185: have profound influences on the development or existence of life
186: (e.g., \citealt{hoffman+schrag2002}), identifying the likelihood of
187: such transitions on terrestrial exoplanets should be central to any
188: habitability assessment.  Following in the footsteps of our first
189: analysis, here we focus on obliquity and consider the influence on
190: habitability of several planetary attributes a priori unknown for
191: exoplanets, such as the efficiency of latitudinal heat transport and
192: the land-ocean distribution.
193: 
194: The remainder of this paper is structured as follows: In
195: \S~\ref{obl_sec:model} we describe the energy balance climate model we
196: use. In \S~\ref{obl_sec:valid} discuss several validation tests in
197: which our model performs well enough to give us some confidence in its
198: behavior for conditions that differ from those found on Earth. In
199: \S~\ref{obl_sec:results} we examine the influence on regional and
200: seasonal habitability of various excursions from Earth-like
201: conditions.  Finally, we conclude in \S~\ref{obl_sec:disc+conc}.
202: 
203: 
204: \section{Model}
205: \label{obl_sec:model}
206: 
207: In order to describe the surface temperature and its evolution on a
208: terrestrial planet, we use a 1-dimensional time-dependent energy
209: balance model based on a diffusion equation for latitudinal heat
210: transport.  This type of model has been used in previous
211: investigations of habitable climates (WK97, SMS08), in modeling
212: Martian climate under changes in forcing
213: \citep{nakamura+tajika2002,nakamura+tajika2003}, and in studies of the
214: Earth's climate (\citealt{north_et_al1981} and references therein).
215: 
216: Our model is based on the following prognostic equation for the
217: planetary surface temperature (as described in SMS08):
218: \begin{equation}
219: \label{obl_eq:diffu eq1}
220: C \frac{\partial T[x,t]}{\partial t} - \frac{\partial}{\partial x}
221: \left( D(1 - x^2) \frac{\partial T[x,t]}{\partial x} \right) + I = S
222: (1 - A).
223: \end{equation}
224: In this equation, $x \equiv \sin \lambda$ is the sine of latitude
225: $\lambda$, $T$ is the temperature, $C$ is the effective heat capacity
226: of the surface layer, $D$ is the diffusion coefficient that determines
227: the efficiency of latitudinal redistribution of heat, $I$ is the
228: infrared emission function (energy sink), $S$ is the diurnally
229: averaged insolation function (energy source) and $A$ is the
230: albedo. Our formalism for insolation on oblique planets follows that
231: of WK97. In the above equation, $C$, $D$, $I$, and $A$ may be
232: functions of $T$, $x$, $t$, and possibly other relevant parameters.
233: 
234: Our prescriptions for the functions $C$, $D$, $I$, and $A$ follow
235: SMS08 and are largely borrowed from WK97 or the existing geophysical
236: literature on similar energy balance models (EBMs). For simplicity and
237: flexibility, many of our models use very simple, physically motivated
238: prescriptions. As described in SMS08, we find that an infrared cooling
239: function of the form
240: \begin{equation}
241: I[T] = \sigma T^4 / (1 + (3/4)\tau_{\rm IR}[T])
242: \label{obl_eq:I}
243: \end{equation}
244: (i.e., a one-zone model combined with a simple Eddington transfer
245: approximation; e.g., \citealt{shu1982}), reproduces the greenhouse
246: effect on Earth reasonably well for our purposes.  Here, $\tau_{\rm
247: IR}$ represents the opacity of the atmosphere to long wavelength
248: infrared radiation.  SMS08 describes three pairs of infrared radiation
249: functions and albedo functions.  In this analysis, we will use two of
250: the three: $(I_2,A_2)$, which gives the closest match to the Earth's
251: annual average temperature distribution, and $(I_3,A_3)$, which uses
252: the standard linearized cooling function of
253: \citet{north+coakley1979}. The $(I,A)$ functions are detailed in
254: Table~\ref{obl_tab:one}.  The albedo functions $A_2$ and $A_3$ are
255: constant and low ($\sim 0.3$) for high temperatures, constant and high
256: ($\sim 0.7$) for low temperatures (to represent the high albedo of
257: snow and ice), and vary smoothly in between.  One other $I$ function
258: that we use is the one proposed in WK97, derived from full
259: radiative-convective calculations, here denoted $I_{\rm WK97}$.  This
260: function includes the detailed influence of $\rm CO_2$ (and
261: implicitely $\rm H_2O$) atmospheric content on radiative cooling.  For
262: $C$, we assume various configurations of land and ocean, in each
263: configuration using the same land, ocean, and ice partial $C$ values
264: as WK97.  Finally, we adopt a diffusion coefficient for latitudinal
265: heat transport $D_{\rm fid} = 5.394\times 10^2 {\rm
266: ~erg~cm^{-2}~s^{-1}~K^{-1}}\times (\Omega_p/\Omega_\Earth)^{-2}$, as
267: described in SMS08, where $\Omega_p$ is the angular spin frequency of
268: the model planet and $\Omega_\Earth$ is that of the Earth. As
269: explained in SMS08, this scaling largely oversimplifies the complexity
270: of atmospheric transport expected for planets at different rotation
271: rates (e.g., \citealt{delgenio_et_al1993,delgenio+zhou1996}).
272: 
273: Equation~(\ref{obl_eq:diffu eq1}) is solved as described in SMS08, on
274: a grid uniformly spaced in latitude ($1.25\degr$ resolution elements,
275: found to be sufficient from convergence tests).  We again choose ``hot
276: start'' ($T\geq 350\rm~K$) initial conditions, to minimize the
277: likelihood that models will undergo a dynamical transition to fully
278: ice-covered (snowball) states from which they cannot recover because
279: of ice-albedo feedback. In that sense, our results on snowball states
280: are conservative.
281: 
282: To summarize, we make the following assumptions in the models
283: presented below:
284: \begin{enumerate}
285:  \item {\it Heating/Cooling} -- The heating and cooling functions are
286:     given by the diurnally averaged insolation from a sun--like
287:     (1~$M_\sun$, 1~$L_\sun$) star, with albedo and insolation
288:     functions described above and in Table~\ref{obl_tab:one}.
289:  \item {\it Latitudinal Heat Transport} -- We test the influence on
290:     climate of three different efficiencies of latitudinal heat
291:     transport, within the diffusion equation approximation: an
292:     Earth-like diffusion coefficient, and diffusion coefficients
293:     scaled down and up by a factor of 9 (which correspond to 8-hour
294:     and 72-hour rotation according to the above $D\propto
295:     {\Omega_p}^{-2}$ scaling).
296:  \item {\it Ocean Coverage} -- We vary both the fraction and the
297:     distribution of ocean coverage.  For ocean fraction, we present a
298:     series of models with Earth--like (30\%:70\%) land:ocean fraction, and
299:     another series of models that represent a desert-world, with a
300:     90\%:10\% land:ocean fraction.  For ocean distribution, we present
301:     models in which there is a uniform distribution (in every latitude
302:     band) of land and ocean, and others in which the land-mass is a
303:     single continent centered on the North Pole, while the rest of the
304:     planet is covered with ocean.
305:  \item {\it Initial Conditions} -- As described in SMS08, the models
306:     all have a hot-start initial condition, with a uniform surface
307:     temperature of at least 350~K, to minimize the chances of ending
308:     up in a globally-frozen snowball state owing solely to the choice
309:     of initial conditions.  Time begins at the Northern winter
310:     solstice.
311: \end{enumerate}
312: 
313: 
314: \section{Model Validation}
315: \label{obl_sec:valid}
316: 
317: In SMS08, we verified that our ``fiducial'' model (70\% ocean;
318: $I_2,A_2$ cooling-albedo functions) at 1~AU predicts temperatures that
319: match the Earth's actual temperature distribution at all latitudes
320: that are not significantly affected by Antarctica (i.e., north of
321: $60\degr$~S or so).  This indicates that the model accounts for the
322: overall (annual) planetary energy balance reasonably well.  Another
323: obvious test is whether the model correctly predicts the monthly
324: energy fluxes that together go into the overall balance.  Because our
325: current investigation tests the influence of obliquity on climate, and
326: obliquity is the primary driver of the Earth's seasons, verifying the
327: seasonal predictions of our model, given Earth-like conditions, is
328: particularly relevant.
329: 
330: The diffusion equation model is a statement of conservation of energy.
331: By definition, after vertical integration for a thin atmosphere with
332: dominant surface processes,
333: \begin{equation}
334: C \frac{\partial T}{\partial t} \equiv \frac{\partial \sigma}{\partial t},
335: \label{eq:def hc}
336: \end{equation}
337: where $\sigma$ is the energy surface density (internal energy per unit
338: surface area on the globe).  The diffusion equation, therefore, says
339: that the rate of change of internal energy at a given point equals the
340: sources of energy (insolation), minus the sinks (infrared radiation),
341: minus whatever energy flows away from the point under consideration.
342: 
343: Figure~\ref{obl_fig:set30heat_cool_23p5} presents a comparison between
344: the annually averaged fluxes of incoming and outgoing radiative energy
345: in the fiducial model with the corresponding fluxes on Earth, taken
346: from NASA's Earth Radiation Budget Experiment (ERBE) in the mid-1980s
347: \citep{barkstrom_et_al1990}.\footnote{The ERBE satellite measured
348: short-wavelength, or incoming, flux as that from 0.2~$\mu$m to
349: 4.5~$\mu$m.  Long-wavelength, or outgoing, flux was defined as all
350: other flux within the bolometric range of the instrument.}  While our
351: model does not capture the full shape of the Earth's cooling and
352: heating functions -- in particular, the annually averaged model
353: heating function is a bit below the Earth's at the poles -- still,
354: both cooling and heating fluxes are within 10\% of the Earth's over
355: most of the planet's surface.
356: 
357: Figure~\ref{obl_fig:set30heat_cool_monthly23} offers an even more
358: compelling validation.  In this figure, each of the 12 panels shows
359: solar (i.e., heating), terrestrial (i.e., cooling), and net (solar
360: minus terrestrial) radiative fluxes as functions of latitude, averaged
361: over one month.  Not only are our annually averaged cooling and
362: heating functions in reasonable agreement with Earth's, as per
363: Figure~\ref{obl_fig:set30heat_cool_23p5}, but furthermore the temporal
364: variability of radiative fluxes in our model is similar to that of
365: Earth.
366: 
367: For example, at the Northern winter solstice (upper left panel of
368: Fig.~\ref{obl_fig:set30heat_cool_monthly23}), the model heating curve
369: closely traces that of the Earth.  It peaks at a somewhat more
370: Southern latitude than the Earth's does, but is within 10\% of the
371: Earth's at all latitudes north of $60\degr$~S.  As the months advance,
372: the concordance between the model heating curve and the Earth's
373: heating curve increases, until there is maximum agreement (within 10\%
374: at all atitudes) at the equinox (``Solstice+03'').  Then, by the next
375: solstice, the curves agree to within 10\% at all latitudes South of
376: roughly $60\degr$~N. In a comparison of the cooling curves, the model
377: shows even greater agreement with the data.  In a majority of months,
378: these two curves are within 10\% of each other at all latitudes.
379: 
380: Interestingly, the month-by-month variations in model heating and
381: model cooling lead to a net heating curve (heating minus cooling) that
382: predicts some detailed features actually seen in the Earth's net
383: heating budget.  Notice, for instance, the slight upward turn of the
384: net heating curves of both the model and the Earth near the North
385: Pole, at and around the Northern winter solstice.  A similar feature
386: is seen in both curves (though with slightly less impressive detailed
387: agreement) near the South Pole, at and around the Southern winter
388: solstice.  These comparisons establish that our climate model exhibits
389: reasonable regional and seasonal variability of not just temperature
390: but also incoming and outgoing radiative energy fluxes.
391: 
392: Another way to consider seasonal variations of heating and cooling
393: fluxes is to look at the globe--average of each with respect to time.
394: Figure~\ref{obl_fig:set30_heat_cool_vs_time} presents a comparison of
395: these fluxes, for our fiducial model and the Earth itself.  The bottom
396: panel of this figure shows the net heating flux as a function of time
397: of year, measured in fraction of a year from the Northern winter
398: solstice.  Earth's net heating flux varies by about 5\% with respect
399: to the cooling flux, while our model's varies somewhat less.  The
400: heating function for the Earth exceeds the cooling function during
401: Northern winter for two main reasons:\footnote{Note that the cooling
402: function, which traces surface temperatures, varies less through the
403: seasonal cycle than the heating function.} First, the nonzero
404: eccentricity ($e\approx 0.0167$) of the Earth's orbit places its
405: perihelion -- which occurs during Northern winter -- approximately
406: 3.4\% closer to the Sun than its aphelion.  This is responsible for
407: $\sim 7\%$ of the annual $\sim 10\%$ annual variation in net heating
408: flux. Our fiducial model, on the other hand, assumes zero
409: eccentricity. Another contributing factor is that the Earth's oceans
410: on average absorb somewhat more insolation than the land, and the
411: Southern hemisphere -- which faces the Sun during Northern winter --
412: has greater ocean coverage than the Northern hemisphere. To within $5
413: \%$, however, the Earth remains in global radiative balance throughout
414: its seasonal cycle.
415: 
416: 
417: So far we have considered the radiative fluxes, but what about
418: diffusive energy flux?  We may combine equation~(\ref{obl_eq:diffu
419: eq1}) with equation~(\ref{eq:def hc}) to produce:
420: \begin{equation}
421: \frac{\partial \sigma}{\partial t} - \frac{\partial}{\partial x} \left\{ D\cos^2\lambda \frac{\partial}{\partial x}T \right\} = \left({\rm sources - sinks}\right)_{\rm energy~per~area},
422: \label{eq:act. eq. sig}
423: \end{equation}
424: where we have substituted $\cos^2 \lambda$ for $(1-x^2)$.  Comparing
425: equation (\ref{eq:act. eq. sig}) to a diffusion equation in spherical
426: coordinates, and accounting for vertical integration, shows that
427: %
428: $F_\lambda = R D \cos\lambda (\partial T/\partial x)$ is the rate of
429: latitudinal energy transport per unit longitudinal length.  The total
430: rate of meridional diffusive heat transport (i.e. energy crossing a
431: given latitude circle per unit time) therefore is
432: \begin{equation}
433: \mathcal{F}_\lambda = 2\pi R \cos\lambda F_\lambda = 2 \pi R^2 D
434: \cos^2\lambda \frac{\partial T}{\partial x}.
435: \label{eq:total flux}
436: \end{equation}
437: 
438: Figure~\ref{obl_fig:set80merid_heat_flux} shows profiles of this
439: diffusive heat transport rate in our fiducial model, at Earth-like
440: $23.5\degr$ obliquity and at extreme $90\degr$ obliquity.  In the
441: Earth-like configuration, heat flows from the equator toward the
442: poles.  In the highly oblique configuration, however, heat flows in
443: the other direction, from the poles to the equator (in an annually
444: averaged sense). For comparison, \citet{williams+pollard2003} present
445: a full general circulation model (GCM) of an Earth-like planet at
446: Earth-like and higher obliquity.  Figure~2 of that paper shows the
447: meridional heat flux within their models for $23.5\degr$ obliquity and
448: $85\degr$ obliquity, and the results are strikingly similar to ours.
449: At $23.5\degr$ obliquity, our model's diffusive flux is very close to
450: that of the GCM.  At high obliquity, the flux in our model remains
451: within $\sim 30\%$ of that in the GCM (from visual inspection), at all
452: latitudes.  This reasonable concordance indicates that the treatment
453: of heat transport within our model, despite being very simple, is
454: still likely to remain useful as a representation of heat transport in
455: less-Earth-like conditions.  We emphasize that it is a nontrivial
456: point that this entirely different regime of transport should remain
457: well-captured by a diffusion approximation.
458: 
459: 
460: 
461: 
462: \section{Study of Habitability}
463: \label{obl_sec:results}
464: 
465: For model planets with $23.5\degr$ obliquity on a circular orbit at
466: 1~AU, both pairs of infrared cooling functions and albedo functions
467: presented in Table~\ref{obl_tab:one} are reasonably good matches for
468: the Earth's current climate, as measured by latitudinally averaged
469: temperatures, with a somehwat better fit with $(I_2,A_2)$.  This gives
470: us some confidence that these functions are useful guides as to how
471: the climate might respond under different forcing conditions.  In this
472: investigation, we consider how variations in intrinsic planetary
473: characteristics combine with the changes in insolation and year-length
474: at various orbital radii to map the zone of regionally habitable
475: climates on planets with various obliquities.
476: 
477: We follow SMS08 in saying that, at a given time, a part of a planet is
478: habitable if its surface temperature is between 273~K and 373~K,
479: corresponding to the freezing and boiling points of pure water at
480: 1~Atm pressure.  This criterion may be criticized for several reasons
481: discussed in SMS08 and references therein, but it provides a
482: reasonable starting point for making numerical investigations.  We
483: will frequently quantify habitability of pseudo-Earths with the
484: temporal habitability fraction, $f_{\rm time}[a,\lambda]$, where $a$
485: is orbital semimajor axis, $\lambda$ is latitude, and $f_{\rm time}$
486: is the fraction of the year that the point in parameter space
487: specified by $(a,\lambda)$ spends in the habitable temperature range
488: (see SMS08 for details).
489: 
490: 
491: \subsection{Efficiency of Heat Transport}
492: \label{obl_ssec:efficiency}
493: 
494: Terrestrial planets with different rotation rates will redistribute
495: heat from the substellar point (or, in a 1D model, the substellar
496: latitude) with different efficiencies.  According to the idealized
497: scaling described above, wherein the effective diffusion coefficient
498: varies with the inverse square of the planetary rotation rate, slower
499: spinning planets will redistribute heat more efficiently, while faster
500: spinning planets will do so less efficiently.  But from where, and to
501: where, is heat redistributed?  How does this depend on obliquity and
502: rotation rate?  And what influence does this have on climatic
503: habitability?
504: 
505: \subsubsection{Direction of Heat Flow}
506: \label{obl_sssec:direction}
507: 
508: For an Earth-like $23.5\degr$ obliquity, the substellar latitude does
509: not vary very much over the course of the year: the tropics are fairly
510: close to the equator (the tropical region is less than one third of
511: the Earth's surface area).  As a result, it is a reasonable
512: approximation that heat is always being transported from the equator
513: to the poles (but see Fig.~\ref{obl_fig:set30heat_cool_monthly23} for
514: details).  In contrast, on a planet with significantly larger
515: obliquity, the direction of heat flow changes over the course of the
516: annual cycle.  At the equinoxes, the equator is the most strongly
517: insolated part of the planet (regardless of obliquity), and so heat
518: builds up at the equator, to be partially redistributed by atmospheric
519: motions.  But on a highly oblique planet, polar summers are extremely
520: intense, as measured by diurnally averaged insolation.  As a result,
521: heat builds up at the poles during their corresponding summers, and
522: the flow of heat reverses direction.
523: 
524: Figure~\ref{obl_fig:set30heat_cool_6090} demonstrates the effect of
525: such strong polar summers on the global radiation budget of a model
526: planet, by comparison to Figure~\ref{obl_fig:set30heat_cool_23p5} in
527: which the annually averaged cooling and heating are shown for an
528: Earth-like, $23.5\degr$ obliquity model.  As expected for the
529: Earth-like model, over the annual cycle, the equator receives
530: significantly more solar radiation than do the poles, and accordingly
531: the annually averaged heating exceeds the cooling at the equator.
532: This indicates that atmospheric motions transport heat poleward from
533: the equator on average.  In Figure~\ref{obl_fig:set30heat_cool_6090},
534: on the other hand, we present the analogous functions in the case of
535: high and extreme obliquity models. The left panel shows the heating
536: and cooling functions for a model at $60\degr$ obliquity; the right
537: panel shows the same functions for a model at $90\degr$ obliquity.
538: Planetary scientists have long recognized that in highly oblique
539: models such as these, the polar summers are so intense that, averaged
540: over the year, the most strongly insolated parts of the planet are the
541: North and South Poles!  (See, e.g., \citealt{ward1974}.)  In an
542: annually averaged sense, then, heat flows from the poles to the
543: equator, although clearly the direction of flow changes with the
544: seasons, as described above.  The import of these plots is that our
545: notion that the poles are the coldest planetary regions might have to
546: be revised in the case of highly oblique worlds. The resulting regime
547: of atmospheric transport, which is only parameterized with our
548: diffusive treatment, may also be expected to differ substantially from
549: that on Earth (e.g., in terms of equatorial Hadley cells).
550: 
551: 
552: %\clearpage
553: 
554: Figure~\ref{obl_fig:set30heat_cool_monthly90} shows in greater detail
555: the extreme way in which insolation can vary over the annual cycle in
556: a highly oblique model.  In this model, the obliquity of an otherwise
557: Earth-like planet ($I_2$--$A_2$ functions, with 70\% ocean uniformly
558: distributed) is set to 90$\degr$.  Notice that the cooling remains
559: much more steady than the heating in this model.  This is because of
560: the high effective heat capacity of the atmosphere above ocean.  In
561: models with less ocean coverage, or oceans that are nonuniformly
562: distributed, the cooling too can vary dramatically over the annual
563: cycle.\footnote{It is also worth noting that at the outer reaches of a
564: $\sim 2M_\sun$ star's habitable zone
565: \citep[e.g.,][]{kasting_et_al1993}, the annual cycle might be long
566: enough relative to the thermal timescale ($\sim$ a decade) of the
567: ocean-atmosphere mixed layer that it could undergo relatively large
568: swings in temperature within a single annum.}  Because the heating
569: varies so intensely while the cooling varies less, their difference --
570: the net heating curve -- also exhibits large variations within the
571: annual cycle.  At each solstice, the pole facing the star receives far
572: more net radiant flux than both the opposite pole and the planetary
573: mean.  At the equinoxes, something perhaps more surprising happens:
574: the net radiant flux is negative over much of the model planet's
575: surface, and only barely positive near the equator.  Overall, the
576: planet heats strongly at the poles during solstices (while cooling
577: elsewhere) and either cools or remains essentially thermally neutral
578: everywhere during equinoxes.
579: 
580: 
581: %\clearpage
582: 
583: 
584: \subsubsection{Implications for Habitability}
585: \label{obl_sssec:hab}
586: 
587: As SMS08 demonstrates, model planets with efficient heat transport
588: (slowly spinning planets) are more latitudinally isothermal than
589: models with Earth-like rotation, which themselves exhibit less
590: latitudinal variation of temperature than those with inefficient heat
591: transport (fast spinning planets).  As a result, models
592: corresponding to worlds that are spinning slowly relative to the Earth
593: (but still fast enough that a 1-D climate model has some use) tend to
594: be either entirely habitable or entirely non-habitable at any given
595: time.  In contrast, Earth-like and faster spinning model planets may
596: be only partially habitable at a particular time.  They may, for
597: instance, be frozen at the poles and temperate at the equator, or vice
598: versa in the case of a highly oblique world.
599: 
600: Figure~\ref{obl_fig:obliq_rotI2A2I3A3} demonstrates the complicated
601: interplay that can go on between obliquity and efficiency of heat
602: transport, in determining a planet's habitability.  This figure shows
603: the temporal habitability fraction, as a function of orbital semimajor
604: axis and latitude, for each of 12 different combinations of obliquity
605: ($0\degr$, $30\degr$, $60\degr$, $90\degr$) and latitudinal heat
606: diffusion coefficient ($D_{\rm fid}/9$, $D_{\rm fid}$, and $9D_{\rm
607: fid}$, corresponding respectively to 72-hour, 24-hour, 8-hour
608: rotations).  The top panel shows these plots for the ($I_2,A_2$)
609: cooling--albedo pair, and the bottom panel depicts the ($I_3,A_3$)
610: pair (see Table~\ref{obl_tab:one} for details).  In both panels, the
611: left column of plots represents efficient latitudinal heat transport;
612: the middle column represents Earth-like transport; and the right
613: column represents inefficient transport.  Each of the 24 plots in this
614: figure shows the results of model runs for planets located from
615: $0.45$~AU to 1.25~AU in increments of $0.025$~AU. The color scale
616: indicates the fraction of the year that the latitude at that point
617: spends in the habitable temperature range (273~K - 373~K) on a model
618: planet at the specified orbital semimajor axis.  In each plot, the
619: white vertical dashed lines indicate the radiative equilibrium
620: habitable zone, calculated (as discussed in SMS08) from a
621: 0-dimensional model with annually averaged, globally averaged
622: insolation and cooling.
623: 
624: There are a number of interesting features in
625: Figure~\ref{obl_fig:obliq_rotI2A2I3A3}.  The most obvious one is that,
626: as expected, at every obliquity, less efficient transport results in
627: more strongly latitudinally differentiated temporal habitability.  In
628: addition, at each transport-efficiency value, the {\tt $>$}~sign shape
629: of the seasonally habitable ribbon at low obliquity reverses to a {\tt
630: $<$}~sign shape at high obliquity.  In other words, at low obliquity,
631: the relatively cold poles are habitable closer to the star and the
632: relatively warm equator is habitable farther from the star. At high
633: obliquity, however, this reverses and the poles are relatively warm,
634: while the equator is comparatively colder.
635: 
636: Furthermore, in both panels, the contours in most cells show a very
637: abrupt outer boundary to the seasonally habitable zone.  This is
638: because, as discussed in SMS08, the ice-albedo feedback renders these
639: models quite sensitive to changes in forcing.  Small reductions in
640: insolation can be amplified, because the ice-coverage increases, which
641: increases the global albedo and leads to further reduction in
642: insolation.  This feedback mechanism renders the model climates
643: susceptible to a global snowball transition, from which they cannot
644: recover within our model framework.  \footnote{Additional feedback
645: mechanisms that are not incorporated in our model exist on a real
646: planet and might help it to recover from a snowball state.  For
647: discussions of how the Earth might have recovered from one or more
648: snowball episodes, see, for example, \citet{caldeira+kasting1992},
649: \citet{hoffman+schrag2002}, and \citet{pierrehumbert2004}.}  The main
650: exceptions to this trend are the low obliquity, fast-spinning models,
651: in the upper right corners of both panels, although even these models
652: drop to 0\% habitability at orbital radii that are small relative to
653: the outer boundary of the habitable zone set by global radiative
654: equilibrium (indicated by the white dashed lines).  Interestingly, the
655: fairly small difference in cooling--albedo functions from ($I_2,A_2$)
656: to ($I_3,A_3$) is sufficient to allow the cell in the lower right
657: corner of the bottom panel -- extreme obliquity, inefficient transport
658: -- to avoid transitioning globally to a snowball state.  In that
659: model, the intense summer insolation at the poles, combined with the
660: relative thermal isolation of different latitudes, allows the poles to
661: heat up above the freezing point of water during their summers, even
662: at orbital distances where other models would be entirely frozen.  In
663: sum, susceptibility to snowball transitions depends on details of
664: parameterizations in our energy balance model, as has been noted
665: before for Earth climate studies \citep[e.g.,][]{north_et_al1981}.
666: 
667: 
668: \subsection{Land/Ocean Distribution}
669: \label{obl_ssec:ocean distribution}
670: 
671: As described in SMS08, the large covering fraction of oceans on the
672: Earth (roughly 70\%) stabilizes our climate over an annual cycle, by
673: virtue of the large effective heat capacity of atmosphere over ocean.
674: Over land, the thermal relaxation timescale is several months, while
675: for the ($\sim 50$~m deep) mixed layer over the ocean, the thermal
676: relaxation timescale is more than a decade.  As a result, in a 1D
677: model (such as ours) that does not resolve continents in longitude,
678: any latitude band with significant ocean fraction will have strongly
679: suppressed annual temperature variations relative to a latitude band
680: with low ocean fraction.  Because we do not know of any way to
681: determine a priori the distribution of continents and oceans on an
682: extrasolar planet, it is important to consider the influence on
683: climatic habitability of other possible land/ocean distributions.
684: 
685: \subsubsection{Nonuniform Ocean Coverage}
686: \label{obl_sssec:nonuniform}
687: 
688: We consider model planets with distributions of land and ocean that
689: are not uniform across different latitudes: one with 30\% land
690: coverage, with a land-mass centered on the North Pole (extending down
691: to $\sim 24\degr$ North latitude), and the other (discussed in
692: \S~\ref{obl_sssec:desert}) with 90\% land, again centered on the North
693: Pole.\footnote{Equivalently, this model can be conceived as having an
694: ocean centered at the South Pole.}  Because of the relatively low
695: thermal inertia of atmosphere over land, parts of a model planet that
696: are dominated by land can freeze or boil during the course of the year
697: and still return to temperate conditions at other times.  In fact, at
698: some orbital distances, and at high obliquity, the polar regions of
699: some models freeze {\it and} boil within an annual cycle.
700: 
701: Figure~\ref{obl_fig:set31_23p5_60_90} displays the tremendous swings
702: of temperature that can occur over latitude bands that lack ocean, and
703: also indicates that annually averaged calculations can miss a lot of
704: information about instantaneous conditions on oblique planets.  This
705: figure contrasts the annually averaged temperature with the detailed
706: temperature evolution on a model planet with a North Polar continent
707: that is 30\% of the total surface area, at an orbital distance of
708: 1~AU.  This model uses ($I_2,A_2$) cooling--albedo functions and
709: results are shown for obliquities $23.5\degr$, $60\degr$, and
710: $90\degr$.  At all three obliquities, the left column -- the annually
711: averaged temperature profile -- provides an impoverished view of the
712: actual climatic conditions.  Looking at just the left panels: the
713: $23.5\degr$ obliquity model appears slightly asymmetrical in
714: temperature distribution, with the continental North Pole 8~K warmer
715: than the oceanic South Pole; the $60\degr$ obliquity model appears
716: cooler at the continental pole; and the $90\degr$ obliquity model
717: again appears warmer at the continental pole, but appears frozen over
718: the whole globe.  In truth, all three models reach significantly
719: higher temperatures at the continental pole during its summer than at
720: the other pole.  At both $60\degr$ and $90\degr$ obliquity, North Pole
721: summer temperatures exceed 410~K, as the Sun shines nearly straight
722: down on the pole for months.  We note that an important limitation of
723: our models is apparent in this figure. Although we may not have much
724: intuition for what the polar summers should be like on high obliquity
725: planets, it is surprising to obtain summer polar continent
726: temperatures in excess of 310~K in the $23.5\degr$ obliquity model.
727: Indeed, Antarctica -- Earth's continental pole -- is significantly
728: {\it colder} than the non-continental pole, and for the most part
729: neither pole ever reaches temperatures above freezing.  Accounting for
730: Antarctica in our model framework would be non-trivial, as it may
731: require initial conditions differing from uniformly ice-free and/or
732: improved treatments of ice-related surface processes.
733: 
734: Figure~\ref{obl_fig:set26_obliq} presents plots of the temporal
735: habitability fraction for the same model planet as above, at
736: obliquities $0\degr$, $30\degr$, $60\degr$, $90\degr$.
737: Figure~\ref{obl_fig:set31_23p5_60_90} illustrated how the presence of
738: land at the North Pole causes tremendous swings in temperature there;
739: Fig.~\ref{obl_fig:set26_obliq} confirms that, for nonzero obliquity,
740: this is indeed the case throughout the orbital extent of the habitable
741: zone.  These large seasonal variations lead to exotic shapes in plots
742: of temporal habitability.  Compared with uniformly ocean-dominated
743: worlds, much more of the parameter space is partially habitable at
744: each obliquity (except $0\degr$), neither 0\% nor 100\% of the year,
745: but somewhere in between.
746: 
747: 
748: \subsubsection{Desert Worlds}
749: \label{obl_sssec:desert}
750: 
751: We now consider two model planets with just 10\% ocean fraction.  We
752: examine the cases of a uniformly distributed ocean (10\% in every
753: latitude band) and an ocean localized at the South Pole (extending
754: northward to $\sim 53\degr$ South latitude).
755: Figure~\ref{obl_fig:set25_obliq} presents the temporal habitability
756: for the uniform desert world, while Figure~\ref{obl_fig:set28_obliq}
757: presents the analogous plot for the desert world with a South Polar
758: ocean.  As before, some regions of these model planets swing from
759: freezing to boiling temperatures over the course of the year.  This is
760: responsible for the butterfly shape of the temporal habitability plots
761: in the $60\degr$ and $90\degr$ obliquity cases shown in
762: Fig.~\ref{obl_fig:set25_obliq}: at $a\sim 0.9\rm~AU$, the poles are
763: habitable for a smaller fraction of the year than more equatorial
764: regions at that orbital distance, or than the poles for closer and
765: more distant orbits. The pattern of habitability in
766: Fig.~\ref{obl_fig:set28_obliq}, on the other hand, shows how strongly
767: assymmetric the climate can be on a desert world with a polar ocean.
768: 
769: These models, and those presented in \S~\ref{obl_sssec:nonuniform},
770: suggest that at extreme obliquity the inner edge of the zone of
771: regionally and seasonally habitable climates can be extended
772: dramatically inward, while the outer boundary can only be extended
773: mildly outward.  Several important caveats should accompany this
774: observation, however.  Assuming an infrared cooling function that is
775: constant with orbital radius probably leads to a flawed treatment at
776: both the high and low insolation limits of these models.  At the inner
777: edge of the habitable zone, large increases in atmospheric water
778: content can cause a reduction in the cooling efficiency, leading to a
779: runaway greenhouse effect. An eventual catastrophic water loss can
780: result in a Venus-like outcome, as described by
781: \citet{kasting_et_al1993} and references therein (although this type
782: of outcome might be mitigated by reduced heating from increased
783: cloud-albedo, as mentioned in SMS08).  At the outer edge, on long
784: timescales, a reduced efficiency of the carbonate-silicate weathering
785: cycle is likely to lead to a significant increase in the partial
786: pressure of atmospheric CO$_2$ \citep{kasting_et_al1993}, which could
787: extend the habitable zone from $\sim 1\rm~AU$, as in our models, to
788: $\sim 1.4\rm~AU$ or beyond in some cases.  We discuss in greater
789: detail the issue of varying atmospheric CO$_2$ content with orbital
790: distance in \S~\ref{obl_ssec:IWK97}.
791: 
792: Notwithstanding these various complications, for plausible cooling
793: functions, the low thermal inertia of atmosphere over land might lead
794: to severe polar climates on highly oblique planets.  What are we to
795: make of partial ``habitability'' by our criterion in the case of a
796: region of a planet that actually boils and freezes every year?  There
797: are some microbes on Earth that can reproduce at freezing
798: temperatures, and others that can reproduce at boiling temperatures
799: (see SMS08 and references therein), although none of which we are
800: aware that can do both.  If part of a planet regularly swings through
801: these wild extremes of climate, is it appropriate to call it
802: habitable?  More relevant from the perspective of formulating a
803: testable scientific hypothesis, could such a planet support enough
804: life to produce sufficient levels of biosignatures that its life could
805: be detected from Earth?  This is an open question, but it is
806: worthwhile to keep in mind that microbes on Earth appear to be as
807: hardy as they need to be: nearly everywhere that biologists have
808: searched, they have found some microbes thriving.  A perhaps equally
809: significant result from the perspective of habitability is that the
810: reduced thermal inertia of these models appears to render them
811: somewhat less susceptible to global snowball events, especially at
812: high obliquities (e.g., compare
813: Figs.~\ref{obl_fig:set26_obliq}-\ref{obl_fig:set28_obliq} with the
814: middle column of Fig.~\ref{obl_fig:obliq_rotI2A2I3A3}).
815: 
816: 
817: 
818: \subsection{Modeling the Far Reaches of the Habitable Zone}
819: \label{obl_ssec:IWK97}
820: 
821: \citet{walker_et_al1981} propose that a planet's temperature is
822: regulated on long timescales by a feedback mechanism involving
823: weathering of silicate rocks through carbonic acid from CO$_2$
824: disolved in water.  They argue that since the rate of weathering (and
825: hence of removal of CO$_2$ from the atmosphere) increases with
826: temperature, this process is an important negative feedback on climate
827: that acts to keep temperatures near the freezing point of water (see
828: the recent study by \citealt{zeebe+caldeira2008} confirming the
829: operation of this cycle).  \citet{kasting_et_al1993} point out that
830: this negative feedback can significantly offset the extreme
831: sensitivity of climate to changes in orbital distance away from 1~AU
832: seen in models such as those of \citet{hart1979} and the models
833: presented in SMS08 and thus far in this paper.  These models are
834: sensitive because they contain a significant positive feedback of the
835: Earth-climate -- the ice-albedo feedback whereby at lower temperatures
836: the absorbed insolation is dramatically reduced because of the high
837: albedo of ice.  However, these models also ignore the long-term,
838: negative (counterbalancing) feedback of the carbonate-silicate
839: weathering cycle \citep{kasting_et_al1993}.
840: 
841: In order to probe the combined influence on climate of rotation rate
842: and obliquity in the context of the expected CO$_2$-rich atmosphere
843: that a pseudo Earth would have at 1.4~AU, we switch to the infrared
844: cooling function used by WK97 ($I_{\rm WK97}$ in
845: Table~\ref{obl_tab:one}), with CO$_2$ partial pressure ($p$CO$_2$) set
846: to 1~bar and 2~bars. For simplicity, we maintain the same albedo
847: function $A_2$ as before and adopt the same linear dependence of the
848: latitudinal heat diffusion coefficient with total atmospheric pressure
849: as WK97 ($D/D_{\rm fid} \propto P_{\rm tot}$).  We find that at both
850: 1- and 2-bar levels of CO$_2$, model planets maintain globally
851: temperate conditions at all obliquities for both $D=9D_{\rm fid}$ and
852: $D=D_{\rm fid}$ -- corresponding to slow and Earth-like rotation.
853: Interestingly, however, at reduced transport efficiency, corresponding
854: to fast planetary rotation, we find that these CO$_2$-rich models are
855: susceptible to global glaciation events.
856: 
857: 
858: Figure~\ref{obl_fig:set45_23p5_60_90} shows the global average
859: temperature and the climate evolution for fast-spinning model planets
860: at 1.4~AU with 1~bar atmospheric CO$_2$, ($I_{\rm WK97},A_2$)
861: cooling-albedo functions and a latitudinal heat diffusion coefficient
862: scaled up with total pressure but reduced by a factor 9 to account for
863: rapid rotation.  At $23.5\degr$ obliquity, the cold temperatures at
864: the poles drag the model into a snowball state, while at $90\degr$
865: obliquity, it is the cold equator that drags the model planet to the
866: same fate. At $60\degr$ obliquity, however, no part of the planet
867: receives consistently low enough insolation to trigger global
868: glaciation.  The snowball effect seen in the $23.5\degr$ and $90\degr$
869: obliquity models might be particularly dramatic because of the
870: possibility of partial atmospheric collapse, on a much shorter
871: timescale than volcanism can replenish CO$_2$.  At 1~bar, the freezing
872: point of CO$_2$ is $\sim 195 \rm~K$.  Both the $23.5\degr$ and
873: $90\degr$ obliquity models reach temperatures below this threshold
874: over large enough regions of their surfaces\footnote{The very large
875: seasonal variations of temperature on these snowball planets result
876: from the moderate heat capacity of the atmosphere over ice (see, e.g.,
877: SMS08; WK97).} that significant amounts of the atmospheric CO$_2$
878: might condense out as dry ice, thereby reducing the atmospheric
879: greenhouse effect.  The risk of atmospheric collapse would be somewhat
880: lessened by the release of latent heat from CO$_2$ condensation, which
881: would tend to prevent too much CO$_2$ from freezing out during any
882: winter. Partial collapse, on the other hand, would reduce surface
883: pressure and thus the efficiency of atmospheric heat
884: transport. Realistically treating these possibilities is beyond the
885: scope of the present paper as it would require incorporating a latent
886: heat term in the energy balance equation and accounting for the
887: surface CO$_2$ mass budget, as \citet{nakamura+tajika2002} do in their
888: Mars EBM.
889: 
890: While model runs with 2~bars of CO$_2$ (not shown) indicate that a
891: pseudo-Earth with such a massive atmosphere at 1.4~AU would be
892: unlikely to suffer glaciation or partial atmospheric collapse at any
893: obliquity, it is worth noting that on a planet with a rate of volcanic
894: greenhouse gas output not much higher than on Earth, it may be
895: difficult to build up a thick CO$_2$ atmosphere (e.g., 2 bars) under
896: conditions such that CO$_2$ condenses out at lower pressure (like in
897: our model with 1 bar of CO$_2$).  In situations in which atmospheric
898: CO$_2$ concentrations can change appreciably on a yearly timescale,
899: our model is not self-consistent.  Further analysis is therefore
900: needed to determine the full extent and consequences of such
901: atmospheric condensation events.  Clearly, these various issues are
902: important subjects for future studies since they indicate that
903: snowball transitions could in principle limit the habitability of some
904: terrestrial planets by interfering with the negative feedback of their
905: carbonate-silicate weathering cycles.
906: 
907: 
908: 
909: \subsection{Validity of Global Radiative Balance}
910: \label{obl_sssec:global_balance}
911: 
912: Calculations of habitable zones have often assumed global radiative
913: balance conditions.  Although these calculations by definition cannot
914: account for the regional character of habitability, one might hope
915: that they still provide a decent proxy for the global average
916: conditions.  Indeed, as seen in
917: Fig.~\ref{obl_fig:set30_heat_cool_vs_time}, the Earth itself is within
918: $\sim 5\%$ of radiative equilibrium throughout the year.  Accordingly,
919: global radiative balance models have provided a very useful starting
920: point for considerations of how the habitability of an Earth-like
921: planet depends on its orbital radius.
922: 
923: Figure~\ref{obl_fig:set31_heat_cool_vs_time} presents globally
924: averaged cooling, heating, and net radiative fluxes in a model with a
925: North Polar continent that covers 30\% of the planet's surface, with
926: $90\degr$ obliquity.  In contrast to the Earth, which remains within
927: $5\%$ of global radiative balance, this model planet can be nearly
928: 60\% out of global radiative balance.  While it has been recognized
929: that planets on highly eccentric orbits experience forcings that are
930: significantly different from annually and globally averaged
931: conditions, the results in Fig.~\ref{obl_fig:set31_heat_cool_vs_time}
932: illustrate how, even on circular orbits, planets can experience
933: conditions that are far from radiative equilibrium.  This further
934: underscores the importance of regional, time-dependent climate models
935: for addressing the habitability of extrasolar terrestrial planets.
936: 
937: 
938: 
939: \section{Conclusions}
940: \label{obl_sec:disc+conc}
941: 
942: We have presented a series of energy balance models to address the
943: variety of climatic conditions that might exist on oblique terrestrial
944: planets with circular orbits.  We considered dynamic climate forcings
945: and responses determined by several planetary attributes a priori
946: unknown for extrasolar planets, including obliquity, rotation rate,
947: distribution of land/ocean coverage, and the detailed nature of the
948: radiative cooling and heating functions.  We find that planets with
949: small ocean fractions or polar continents can experience very severe
950: seasonal climatic variations, but that these planets also might
951: maintain seasonally and regionally habitable conditions over a larger
952: range of orbital radii than more Earth-like planets.  Climates on high
953: obliquity planets with nonuniform distributions of land and ocean can
954: be far from global radiative balance, as compared to the Earth. Our
955: results provide indications that the modeled climates are somewhat
956: less prone to dynamical snowball transitions at high obliquity.  Fast
957: rotating Earth-like planets may fall victim to global glaciation
958: events at closer orbital radii than slower rotating planets. This is
959: also the case for planets with massive CO$_2$ atmospheres, which are
960: expected to be found in the outer orbital range of habitable zones.
961: Snowball transitions could be particularly significant for such
962: planets since partial collapse of their CO$_2$-rich atmospheres may
963: occur and possibly interfere with the thermostatic effect of their
964: carbonate-silicate weathering cycle, thus affecting their long-term
965: habitability.
966: 
967: \section*{Acknowledgments}
968: We acknowledge helpful conversations with James Cho, Michael Allison,
969: Anthony Del Genio, and Scott Gaudi.  We thank Diana Spiegel for help
970: with ERBE data.  We acknowledge many useful comments by an anonymous
971: referee. CS acknowledges the funding support of the Columbia
972: Astrobiology Center through Columbia University's Initiatives in
973: Science and Engineering, and a NASA Astrobiology: Exobiology and
974: Evolutionary Biology; and Planetary Protection Research grant, \#
975: NNG05GO79G.
976: 
977: 
978: %\clearpage
979: 
980: \clearpage
981: 
982: \begin{table}[p]
983: \begin{center}
984: \caption[Atmospheric Models]{~~Atmospheric Models.}
985: \vspace{0.2in}
986: \begin{tabular}{cll}
987:   \tableline
988:   \tableline
989:   Model             & IR Cooling Function      & Albedo Function\\[0.1in]
990:   \tableline
991: 2\tablenotemark{a}  &  $I_2[T]  =  \frac{\sigma T^4}{1 + (3/4)\tau_{\rm IR}[T]}$   &  $A_2[T] =  0.525 - 0.245 \tanh[\frac{(T-268{\rm K})}{5{\rm K}}]$   \\
992: 3\tablenotemark{b}  &   $I_3[T]  =  A + B T$   &  $A_3[T] =  0.475 - 0.225 \tanh[\frac{(T-268{\rm K})}{5{\rm K}}]$  \\
993: 4\tablenotemark{c}  &   $I_{\rm WK97}[T,p\rm{CO_2}]$   &  $A_2[T]$  \\
994: \label{obl_tab:one}
995: \end{tabular}
996: \vspace{-0.4cm}
997: \tablenotetext{a}{\,Model with $T$--dependent optical thickness: $\tau_{\rm IR}[T] = 0.79(T/273{\rm K})^3$.}
998: \tablenotetext{b}{\,Linearized model: $A = 2.033\times 10^5{\rm~erg~cm^{-2}~s^{-1}}$, $B = 2.094\times 10^3 {\rm~erg~cm^{-2}~s^{-1}~K^{-1}}$.}
999: \tablenotetext{c}{\,WK97 cooling function (with $A_2$ albedo). The detailed functional form is presented in the appendix of WK97.}
1000: \tablecomments{$\sigma$ is the Stefan-Boltzmann constant.}
1001: \end{center}
1002: \end{table}
1003: 
1004: \clearpage
1005: 
1006: \begin{figure}[p]
1007: \plotone
1008: {fig1.pdf}
1009: \caption[Annually averaged cooling and heating fluxes for fiducial
1010: model at 1~AU and Earth.]{Annually averaged cooling and heating fluxes
1011: for our fiducial model at 1~AU and for the Earth. The thick red line
1012: is infrared cooling and the thick blue line is absorbed solar flux, in
1013: our fiducial model (70\% uniform ocean, $I_2,A_2$). The thin dashed
1014: magenta line is the Earth's annually averaged long wavelength infrared
1015: radiation, and the thin dashed cyan line is the annually averaged
1016: absorbed solar flux measured on Earth. Earth-specific features are not
1017: captured by our symmetric and uniform model.}
1018: \label{obl_fig:set30heat_cool_23p5}
1019: \end{figure}
1020: 
1021: \begin{figure}[p]
1022: \plotone
1023: {fig2.pdf}
1024: \caption[Monthly cooling, heating, and net fluxes for fiducial model
1025: at $23.5\degr$ obliquity, at 1~AU, and for Earth.]{Monthly cooling,
1026: heating, and net fluxes for our fiducial model at $23.5\degr$
1027: obliquity and 1~AU orbital distance, and for the Earth.  Each panel
1028: presents the average cooling, heating, and net (heating minus cooling)
1029: radiative fluxes, as functions of latitude, for one month of the year,
1030: starting at the Northern winter solstice (upper left panel), and
1031: incrementing by one month with each panel to the right.  These fluxes
1032: are presented for both the model (thick solid lines) and the Earth
1033: (thin dashed lines).  Model heating is blue; model cooling is red;
1034: model net heating is green.  Earth heating is cyan; Earth cooling is
1035: deep magenta; Earth net heating is black. Our model captures
1036: reasonably well the seasonal variations of these fluxes.}
1037: \label{obl_fig:set30heat_cool_monthly23}
1038: \end{figure}
1039: 
1040: \begin{figure}[p]
1041: \plotone
1042: {fig3.pdf}
1043: \caption[Global average cooling, heating, and net radiative flux, as
1044: functions of time in fiducial model at $23.5\degr$ obliquity.]{Global
1045: average cooling, heating, and net radiative flux, as functions of time
1046: in our fiducial model at $23.5\degr$ obliquity (solid lines). Earth's
1047: ERBE data are shown as stars. {\it Top Panel:} Global average cooling
1048: (red curves and magenta stars) and heating (blue curves and cyan
1049: stars) fluxes as a function of time of year, measured in fraction of a
1050: year from the Northern winter solstice.  {\it Bottom Panel:} Net
1051: heating flux (heating minus cooling) for the model (green curve) and
1052: ERBE data (black stars), plotted as percent of the corresponding
1053: cooling flux. The Earth remains within $5 \%$ of global radiative
1054: balance throughout its seasonal cycle.}
1055: \label{obl_fig:set30_heat_cool_vs_time}
1056: \end{figure}
1057: 
1058: %%%%%%%%%%%%%% Merid Heat Flux
1059: \begin{figure}[p]
1060: \plotone
1061: {fig4.pdf}
1062: \caption[Annually averaged, longitudinally integrated, meridional heat
1063: transport at $23.5\degr$ and $90\degr$ obliquity.]{Annually averaged,
1064: longitudinally integrated, meridional heat transport rates in models
1065: at $23.5\degr$ and $90\degr$ obliquity.  Transport is positive
1066: northward.  In the Earth-like case (blue solid curve), heat flows from
1067: the equator to the poles; in the highly oblique case (green dashed
1068: curve), the annually averaged heat flow is reduced and in the opposite
1069: direction. This result is in close agreement with a comparable one
1070: obtained by \citet{williams+pollard2003} with a full physics climate
1071: model.}
1072: \label{obl_fig:set80merid_heat_flux}
1073: \end{figure}
1074: 
1075: \begin{figure}[p]
1076: \plottwo
1077: {fig5a.pdf}
1078: {fig5b.pdf}
1079: \caption[Annually averaged cooling and heating fluxes for high and
1080: extreme obliquity model planets model and for Earth.]{Annually
1081: averaged cooling and heating fluxes for high and extreme obliquity
1082: model planets, compared to Earth.  By contrast with
1083: Fig.~\ref{obl_fig:set30heat_cool_23p5}, which shows the annually
1084: averaged heating and cooling for $23.5\degr$ obliquity, here we
1085: present these functions for Earth-like models at $60\degr$ obliquity
1086: (left panel) and at $90\degr$ obliquity (right panel). Thick lines are
1087: model results; thin dashed lines are Earth's ERBE data.  In both high
1088: obliquity cases, unlike on Earth, there is net annually averaged
1089: heating at the poles (i.e., heating exceeds cooling), and, especially
1090: for the extreme obliquity case, net annually averaged cooling at the
1091: equator.}
1092: \label{obl_fig:set30heat_cool_6090}
1093: \end{figure}
1094: 
1095: \begin{figure}[p]
1096: \plotone
1097: {fig6.pdf}
1098: \caption[Monthly cooling, heating, and net fluxes for fiducial model
1099: at $90\degr$ obliquity, at 1~AU, and for Earth.]{Monthly cooling,
1100: heating, and net heating fluxes for the fiducial model at $90\degr$
1101: obliquity and 1~AU orbital distance, and for the Earth.  Each panel
1102: presents the average cooling, heating, and net (heating minus cooling)
1103: radiative fluxes, as functions of latitude, for one month of the year,
1104: starting at the Northern winter solstice (upper left panel), and
1105: incrementing by one month with each panel to the right.  These fluxes
1106: are shown for both the oblique model planet (thick solid lines) and
1107: the Earth (thin dashed lines).  Model heating is blue; model cooling
1108: is red; model net heating is green.  Earth heating is cyan; Earth
1109: cooling is deep magenta; Earth net heating is black. While cooling
1110: fluxes remain relatively steady and uniform on such an oblique planet,
1111: heating and net heating fluxes are subject to very large seasonal
1112: variations.}
1113: \label{obl_fig:set30heat_cool_monthly90}
1114: \end{figure}
1115: 
1116: \begin{figure}[p]
1117: \plotone
1118: {fig7a.pdf}\\
1119: \plotone
1120: {fig7b.pdf}
1121: \caption[Model temporal habitability fraction under different
1122: obliquities, rotation rates, and cooling/heating
1123: functions.]{\footnotesize Temporal habitability fraction in models
1124: with different obliquities, heat transport efficiencies, and
1125: cooling-albedo functions. In both panels, obliquity varies from
1126: $0\degr$ (top row) to $90\degr$ (bottom row). The latitudinal heat
1127: diffusion coefficient varies from $D_{\rm fid}/9$ (left column) to
1128: $D_{\rm fid}$ (center) and $9D_{\rm fid}$ (right).  In each panel, the
1129: abscissa is orbital radius, in AU, and the ordinate is latitude.
1130: Colors indicate the fraction of the year spent by that region in the
1131: habitable temperature range (273~K - 373~K).  {\it Top Panel:}
1132: ($I_2,A_2$) cooling-albedo combination. {\it Bottom Panel:}
1133: ($I_3,A_3$) cooling-albedo combination. For comparison, vertical
1134: dashed lines show the habitable zone extent expected on the basis of
1135: global radiative balance. Regionality and seasonality can extend the
1136: inner reach of the instantaneous habitable zone while often reducing
1137: its outer reach, when global snowball transitions occur.}
1138: \label{obl_fig:obliq_rotI2A2I3A3}
1139: \end{figure}
1140: 
1141: \begin{figure}[p]
1142: \plottwo
1143: {fig8a.pdf}
1144: {fig8b.pdf}\\
1145: \plottwo
1146: {fig8c.pdf}
1147: {fig8d.pdf}\\
1148: \plottwo
1149: {fig8e.pdf}
1150: {fig8f.pdf}\\
1151: \caption[Annually averaged and space-time plot of temperatures on
1152: models with a North Polar continent that takes up 30\% of surface area
1153: at 1~AU, cooling/heating = $I_2,A_2$, obliquity = $23.5\degr$,
1154: $60\degr$, $90\degr$.]{\footnotesize Annually averaged and detailed
1155: time-dependent temperature profiles in models with a North Polar
1156: continent that takes up 30\% of surface area (the other 70\% is ocean)
1157: at 1~AU orbital distance, for ($I_2,A_2$) cooling-albedo functions and
1158: obliquities of $23.5\degr$, $60\degr$ and $90\degr$. {\it Left:} The
1159: magenta curve shows the annually averaged temperature profile for
1160: models with a North Polar continent that extends down to $\sim
1161: 24\degr$ North latitude.  The solid blue and dashed green curves are
1162: for reference -- blue: the fiducial model (identical to this one,
1163: except for 70\% ocean uniformly distributed in every latitude band);
1164: dashed green: the Earth's actual temperature profile, as measured by
1165: NCEP/NCAR in 2004 \citep{kistler_et_al1999, kalnay_et_al1996}.  {\it
1166: Right:} Detailed temperature evolution (latitude vs. time) in the
1167: polar continent models from years 8 through 50. {\it Top Row:
1168: $23.5\degr$.}  {\it Middle Row: $60\degr$.}  {\it Bottom Row:
1169: $90\degr$.} Annually averaged profiles (left) miss much of the
1170: seasonal variations shown by the detailed profiles (right). The
1171: $90\degr$ obliquity model experiences an asymmetric, partial snowball
1172: transition.
1173: 
1174: }
1175: \label{obl_fig:set31_23p5_60_90}
1176: \end{figure}
1177: 
1178: \begin{figure}[p]
1179: \plotone
1180: {fig9.pdf}
1181: \caption[Model temporal habitability fraction under different
1182: obliquities, North Polar continent covering 30\% of surface.]{Temporal
1183: habitability fraction at different obliquities for a model planet with
1184: a North Polar continent covering 30\% of its surface. Obliquity varies
1185: from $0\degr$ to $90\degr$ from top to bottom. The latitudinal heat
1186: diffusion coefficient is kept to its fiducial value.  The notation is
1187: similar to Figure~\ref{obl_fig:obliq_rotI2A2I3A3}: colors indicate the
1188: fraction of the year spent by that region in the habitable temperature
1189: range (273~K - 373~K). Habitability has a strong regional and seasonal
1190: character on such a planet, when oblique.}
1191: \label{obl_fig:set26_obliq}
1192: \end{figure}
1193: 
1194: \begin{figure}[p]
1195: \plotone
1196: {fig10.pdf}
1197: \caption[Model temporal habitability fraction under different
1198: obliquities, 10\% ocean uniformly distributed.]{Temporal habitability
1199: fraction at different obliquities for a model planet with 10\% ocean
1200: uniformly distributed.  Otherwise similar to
1201: Figure~\ref{obl_fig:set26_obliq}. Habitability has a strong regional
1202: and seasonal character on such a planet, when oblique.}
1203: \label{obl_fig:set25_obliq}
1204: \end{figure}
1205: 
1206: \begin{figure}[p]
1207: \plotone
1208: {fig11.pdf}
1209: \caption[Model temporal habitability fraction under different
1210: obliquities, North Polar continent covering 90\% of surface (South
1211: Polar ocean).]{Temporal habitability fraction at different obliquities
1212: for a model planet with a North Polar continent covering 90\% of its
1213: surface (i.e., a localized South Polar ocean).  Otherwise similar to
1214: Figure~\ref{obl_fig:set26_obliq}. Habitability has a strong regional
1215: and seasonal character on such a planet, when oblique.}
1216: \label{obl_fig:set28_obliq}
1217: \end{figure}
1218: 
1219: %%%%%%%%%%%%%% Set45
1220: \begin{figure}[p]
1221: \plottwo
1222: {fig12a.pdf}
1223: {fig12b.pdf}\\
1224: \plottwo
1225: {fig12c.pdf}
1226: {fig12d.pdf}\\
1227: \plottwo
1228: {fig12e.pdf}
1229: {fig12f.pdf}
1230: \caption[Annually averaged and space-time plot of temperatures on fast
1231: spinning world at 1.4~AU, WK97 cooling function, $p\rm CO_2 = 1~bars$,
1232: obliquity = $23.5\degr$, $60\degr$, $90\degr$.]{Annually averaged and
1233: detailed time-dependent temperatures on a fast spinning model planet
1234: at 1.4~AU, using the infrared cooling function of WK97 for a
1235: CO$_2$-rich atmosphere with $p\rm CO_2 = 1~bar$. From top to bottom,
1236: results are shown for an obliquity of $23.5\degr$, $60\degr$ and
1237: $90\degr$.
1238: %
1239: The notation and the models are the same as in
1240: Figure~\ref{obl_fig:set31_23p5_60_90}, except for the cooling function
1241: ($I_{\rm WK97}$ with $p\rm CO_2 = 1~bar$) and a latitudinal heat
1242: diffusion coefficient adjusted for fast spin and a massive
1243: atmosphere. In the two globally-frozen models (top and bottom),
1244: surface temperatures that would result in seasonal CO$_2$ atmospheric
1245: collapse ($T \lsim 195$~K) are reached over a fair fraction of the
1246: planet's surface area.}
1247: \label{obl_fig:set45_23p5_60_90}
1248: \end{figure}
1249: 
1250: \begin{figure}[p]
1251: \plotone
1252: {fig13.pdf}
1253: \caption[Global average cooling, heating, and net radiative flux, as
1254: functions of time in model with North Polar continent at $60\degr$ and
1255: $90\degr$ obliquity.]{Global average cooling, heating, and net
1256: (heating $-$ cooling) radiative flux, as functions of time, for a
1257: model planet with a North Polar continent covering $30 \%$ of its
1258: surface, at $90\degr$ obliquity.
1259: %
1260: Notation is similar to
1261: Figure~\ref{obl_fig:set30_heat_cool_vs_time}. This planet experiences
1262: large deviations from global radiative balance.}
1263: \label{obl_fig:set31_heat_cool_vs_time}
1264: \end{figure}
1265: 
1266: 
1267: 
1268: \clearpage
1269: \bibliography{biblio.bib}
1270: 
1271: 
1272: \end{document}
1273: