1: \documentclass[a4paper,german,12pt]{elsart}
2:
3:
4:
5: \setlength{\topskip}{1.cm} \setlength{\footskip}{1.5cm}
6: \setlength{\parindent}{0cm} \setlength{\textheight}{24cm}
7: \setlength{\headsep}{1cm} \setlength{\oddsidemargin}{1.5cm}
8: \setlength{\textwidth}{14.cm} \setlength{\evensidemargin}{1.5cm}
9: \setlength{\topmargin}{0cm} \setlength{\headheight}{0cm}
10:
11: \usepackage{amsmath}
12: \usepackage{amssymb}
13: \usepackage{wasysym}
14: \usepackage{bibentry}
15: \usepackage{epsfig}
16: \usepackage{epsf}
17: \usepackage[dvips]{color}
18: \usepackage{graphicx,graphics}
19: \usepackage{float}
20: \usepackage{natbib}
21: \usepackage{setspace,lineno}
22: \setstretch{1.5}
23:
24: \begin{document}
25: \linenumbers
26:
27: \begin{frontmatter}
28:
29: \title{Warming the early Earth - CO$_2$ reconsidered}
30:
31:
32: \author[DLR]{\corauthref{cor}Philip von Paris},
33: \corauth[cor]{Corresponding author: philip.vonparis@dlr.de}
34: \author[DLR,TU]{Heike Rauer},
35: % change affil. Grenfell PROOF
36: \author[DLR,TU]{J. Lee Grenfell},
37: % change order authors PROOF
38: \author[TU]{Beate Patzer},
39: \author[DLR]{Pascal Hedelt},
40: \author[DLR]{Barbara Stracke},
41: \author[IWF]{Thomas Trautmann} and
42: \author[IWF]{Franz Schreier}
43:
44:
45:
46: \address[DLR]{Institut f\"{u}r Planetenforschung, Deutsches Zentrum f\"{u}r Luft- und Raumfahrt, Rutherfordstr.
47: 2, 12489 Berlin (Germany)}
48:
49: \address[TU]{Zentrum f\"{u}r Astronomie \& Astrophysik, Technische Universit\"{a}t Berlin, Hardenbergstr.
50: 36, 10623 Berlin, (Germany)}
51:
52: \address[IWF]{Institut f\"{u}r Methodik der Fernerkundung, Deutsches Zentrum f\"{u}r Luft- und Raumfahrt, M\"{u}nchener Str. 20, 82234 Wessling (Germany)}
53:
54:
55: \begin{abstract}
56:
57: Despite a fainter Sun, the surface of the early Earth was mostly
58: ice-free. Proposed solutions to this so-called "faint young Sun
59: problem" have usually involved higher amounts of greenhouse gases
60: than present in the modern-day atmosphere. However, geological
61: evidence seemed to indicate that the atmospheric CO$_2$
62: concentrations during the Archaean and Proterozoic were far too low
63: to keep the surface from freezing. With a radiative-convective model
64: including new, updated thermal absorption coefficients, we found
65: that the amount of CO$_2$ necessary to obtain 273 K at the surface
66: is reduced up to an order of magnitude compared to previous
67: studies. For the late Archaean and early Proterozoic period of the
68: Earth, we calculate that CO$_2$ partial pressures of only about 2.9
69: mb are required to keep its surface from freezing which is
70: compatible with the amount inferred from sediment studies. This
71: conclusion was not significantly changed when we varied model
72: parameters such as relative humidity or surface albedo, obtaining
73: CO$_2$ partial pressures for the late Archaean between 1.5 and 5.5
74: mb. Thus, the contradiction between sediment data and model results
75: disappears % add this sentence PROOF
76: for the late Archaean and early Proterozoic.
77:
78: \end{abstract}
79:
80: \begin{keyword}
81:
82: Faint Young Sun problem, Earth - Atmospheres, composition -
83: Radiative transfer
84:
85: \end{keyword}
86:
87: \end{frontmatter}
88:
89: \section{Introduction}
90:
91:
92: Geological evidence has shown that liquid water was present on the
93: Earth's surface earlier than 3.7 Gy ago (e.g.,
94: \citealp{moj1996,rosing2004}) which implies average temperatures on
95: the surface above 273 K. Some authors have even argued for a hot
96: Archaean climate ($T$ $>$340 K), based on oxygen \citep{knauth2003}
97: and silicon \citep{robert2006} isotope analysis of seawater cherts.
98: However, as pointed out by, e.g., \citet{kasthoward2006} and
99: \citet{shields2007}, these isotopic signatures changes might not
100: only be caused by temperature effects. \citet{sleep2006}, for
101: example, deduced a more moderate surface temperature below 300 K,
102: based on quartz weathering records in paleosols. Nevertheless, it is
103: generally accepted that the Earth has been ice-free throughout most
104: of its history.
105:
106: Observations of several solar-type stars of different ages and
107: virtually all standard models of the solar interior have shown that
108: the total solar luminosity has increased since the ZAMS (Zero Age
109: Main Sequence) by about 30\% \citep{gough1981,caldeira1992}. Had the
110: composition of the Earth's atmosphere been the same then as today,
111: the reduced solar flux would have resulted in surface temperatures
112: below 273 K prior to 2.0 Gy \citep{sagan1972}. This apparent
113: contradiction between solar evolution models, climatic simulations
114: and geological evidence for liquid water and moderately warm
115: temperatures on Earth has been termed the "faint young Sun problem".
116:
117: %\newline
118:
119: Numerous studies have attempted to solve this problem. For example,
120: \citet{minton2007} explored the hot early Sun scenario for a
121: non-standard solar evolution. \citet{shaviv2003} showed that
122: moderate greenhouse warming in combination with the influence of
123: solar wind and cosmic rays on climate could resolve the problem.
124: \citet{jenkins2000} assumed high obliquities in a General
125: Circulation Model to account for high Archaean temperatures.
126: %\newline
127:
128: However, the most accepted scenario involves a much enhanced
129: greenhouse effect (GHE) on the early Earth compared to modern Earth.
130: Today, the GHE produces around 30 K of warming, raising the mean
131: surface temperature of the Earth to about 288 K. Increased
132: abundances of greenhouse gases such as carbon dioxide
133: \citep{kasting1987}, methane \citep{pavlov2000}, ethane
134: \citep{haqq2008} or ammonia \citep{sagan1972,sagan1997} will
135: strengthen the GHE, hence potentially resolve the problem. However,
136: all of these studies faced some form of contradictions or large
137: uncertainties, either from geological data on atmospheric conditions
138: or from atmospheric modeling. The formation and destruction of
139: ammonia is highly dependent on UV levels in the atmosphere
140: \citep{sagan1997,pavlov2001}. The hydrocarbon haze necessary to
141: allow higher hydrocarbons to accumulate in the atmosphere depends
142: critically on the CO$_2$/CH$_4$ ratio \citep{pavlov2003}. The high
143: values of methane required to heat the surface of the early Earth
144: depend on estimates of the early biosphere and volcanic activity,
145: which is not well determined. Past CO$_2$ concentrations required by
146: atmospheric models \citep{kasting1987} to reach surface temperatures
147: above 273 K are in conflict with inferred concentrations from the
148: sediment data \citep{hessler2004,rye1995}.
149: %\newline
150:
151: In this work, the role of CO$_2$ in warming the early Earth is
152: reconsidered. We used a one-dimensional radiative-convective model,
153: including updated absorption coefficients in its radiation scheme
154: for thermal radiation. The model was applied to the atmosphere of
155: the early Earth in order to investigate the effect of enhanced
156: carbon dioxide on its climate. Additionally, we investigated the
157: effect of two important parameters, namely the surface albedo and
158: the relative humidity, upon the resulting surface temperature.
159:
160: Our results imply that the amount of CO$_2$ needed to warm the
161: surface of early Earth might have been over-estimated by previous
162: studies. Furthermore, the results show that the contradiction
163: between modelled CO$_2$ concentrations and measured values might
164: disappear by the end of the Archaean.
165:
166: Section \ref{describemodel} describes the model and section
167: \ref{validmracsubsection} the runs to validate the new radiation
168: scheme. The runs performed for this work are explained and
169: summarized in section \ref{about}. In section \ref{showresults}, the
170: results are presented and discussed. Section \ref{conclusions} gives
171: the summary of the results.
172:
173:
174: \section{Atmospheric Model}
175:
176: \label{describemodel}
177:
178: We used a one-dimensional radiative-convective model based on the
179: climate part of the model used by \citet{Seg2003} and
180: \citet{Grenf2007a,Grenf2007b}. Our model differs in upgrades of the
181: radiation scheme to calculate the thermal emission in the
182: atmosphere. The model calculates globally, diurnally-averaged
183: atmospheric temperature and water profiles for cloud-free
184: conditions. We will first state some basic characteristics of the
185: model (\ref{basic}). Then we will describe the calculation of the
186: energy transport via radiative transfer (solar and thermal fluxes)
187: and convection (\ref{tempprofcalc}) to obtain the atmospheric
188: temperature profile. Thereafter, a description of the determination
189: of the water profile is given (\ref{watcalcu}). Finally, the model
190: input parameters are summarized (\ref{summarymodelinput}).
191:
192: \subsection{Basic model description}
193:
194: \label{basic}
195:
196: The model determines the temperature profile by assuming two
197: dominant mechanisms of energy transport, i.e. radiative transfer and
198: convection. The convective lapse rate is assumed to be adiabatic.
199: The radiative lapse rate is calculated from contributions of both
200: solar and thermal radiation, including Rayleigh scattering for solar
201: radiation and continuum absorption in the thermal region. Table
202: \ref{thermalinput} summarizes the contributions of the different
203: atmospheric species to the calculation of the temperature profile.
204:
205: %TABLE \ref{thermalinput}
206:
207: The species considered in the model are molecular nitrogen, water,
208: molecular oxygen, argon, carbon dioxide and carbon monoxide. For
209: example, a typical early Earth run considered 0.77 bar of nitrogen
210: and several (2.9 - 57.2) mb of carbon dioxide in addition to water
211: in varying concentrations (0.5---1 \% at the surface). Molecular
212: nitrogen is an effective Rayleigh scatterer (although not as
213: effective as carbon dioxide) and as a main constituent of the
214: atmosphere also contributes to the heat capacity. Water is not
215: considered to be an important Rayleigh scatterer, but it is relevant
216: for the other radiative processes. Also, water influences the
217: adiabatic lapse rate because it readily condenses in the
218: troposphere. However, due to small mixing ratios, especially in the
219: stratosphere, water vapour does not contribute to the heat capacity
220: of the atmosphere. Molecular oxygen and argon do contribute to the
221: heat capacity of the atmosphere, and molecular oxygen additionally
222: contributes to the Rayleigh scattering coefficient. Carbon dioxide
223: contributes to all relevant radiative mechanisms (molecular
224: absorption of solar and thermal radiation, continuum absorption,
225: Rayleigh scattering), but not to the adiabatic lapse rate because it
226: does not condense under conditions described in this paper. Carbon
227: monoxide is an important absorber species in some mid-infrared
228: windows and contributes to the heat capacity.
229:
230: The model assumes the hydrostatic relation between pressure $p$ and
231: density $\varrho$ throughout the plane-parallel atmosphere. On the
232: 52 model layers, a logarithmic pressure grid is calculated.
233: Specified pressure levels at the planetary surface (e.g., 1 bar for
234: the standard Earth case) and the upper model lid (at
235: 6.6$\cdot10^{-5}$ bar) determine the altitude range which, for
236: modern Earth conditions, extends to 65-70 km, i.e. the lower- to mid
237: mesosphere. For all gases except water, the ideal gas law is taken
238: as the equation of state (see \ref{convadjust} for water). The
239: effect of clouds is difficult to incorporate into 1D models (see,
240: for example, \citealp{pavlov2000,Seg2003}). In the present model,
241: following the approach of \citet{Seg2003}, clouds are implicitly
242: included by adjusting the surface albedo $A_{\rm{surf}}$ such that
243: under modern Earth control conditions the model calculates a mean
244: surface temperature of 288 K. The required value for $A_{\rm{surf}}$
245: is about 0.21, whereas the actual global value for Earth is
246: approximately 0.15. This can be interpreted as a ground cloud layer
247: instead of tropospheric or stratospheric clouds. The model uses a
248: time-stepping algorithm to convergence to the steady-state
249: solution\citep{pavlov2000}.
250:
251: \subsection{Temperature profile}
252: \label{tempprofcalc}
253:
254: During each time step, the temperature profile is calculated from
255: the radiative equilibrium condition. The temperature $T$ on an
256: atmospheric level $z$ is determined by the following equation of
257: energy conservation \citep{pavlov2000}:
258:
259: \begin{equation}\label{timestep}
260: \frac{d}{dt}T(z)=-\frac{g}{c_p(T,z)}\frac{dF(z)}{dp(z)}
261: \end{equation}
262:
263: where $dt$ is the time step in the model, $c_p$ the heat capacity,
264: $F$ the total net radiative flux and $p$ the pressure of the level.
265: The radiative flux $F$ is the sum of thermal planetary and
266: atmospheric emission, $F_{\rm{thermal}}$, and the solar radiative
267: input, $F_{\rm{solar}}$, into the atmosphere:
268:
269: \begin{equation}
270: F(z)=F_{\rm{thermal}}(z)+F_{\rm{solar}}(z) \label{fluesse}
271: \end{equation}
272:
273: These fluxes are calculated separately by two numerical schemes
274: which solve the monochromatic radiative transfer equation (RTE) for
275: the spectral intensity $I_{\nu}$ in the respective spectral domain
276: (i.e., near-UV to near-IR for solar flux, near-IR to far-IR for
277: thermal flux):
278:
279: \begin{equation}
280: \mu\frac{dI_{\nu}}{d\tau_{\nu}}= I_{\nu}-S_{\nu}\label{endgueltig}
281: \end{equation}
282:
283: where $S_{\nu}$ is the source function (either the incident solar
284: radiation or the thermal blackbody emission of the atmospheric
285: layers and the planetary surface), $d\tau_{\nu}$ the optical depth
286: and $\mu=\cos(\theta)$ the cosine of the polar angle. The optical
287: depth is defined as usual by
288:
289: \begin{equation}\label{opttiefe}
290: d\tau=-(\kappa_{\nu}+s_{\nu})dz
291: \end{equation}
292:
293: where $\kappa_{\nu}$ and $s_{\nu}$ represent the absorption
294: coefficient and the scattering coefficient respectively. The
295: absorption coefficient for a gas mixture is calculated from the
296: individual absorption coefficients of the gas species $i$:
297:
298: \begin{equation}\label{kappadefinition}
299: \kappa_{\nu}=\sum_i \kappa_{\nu,i}=\sum_i \sigma_{\rm{abs},i} \cdot N_i
300: \end{equation}
301:
302: where $N_i$ is the number density and $\sigma_{abs,i}$ the molecular
303: absorption cross section of the gas species $i$. When no scattering
304: occurs (i.e., $s_{\nu}=0$), eq. \ref{opttiefe} can be written in
305: terms of the column density $W_i$ of the gas species $i$ as:
306:
307: \begin{equation}\label{optcolumn}
308: \tau=\sum_i \sigma_{\rm{abs},i}\cdot W_i
309: \end{equation}
310:
311: The absorption cross section is defined by:
312:
313: \begin{equation}\label{defabs}
314: \sigma_{\rm{abs}}(\nu,p,T)=\sum_j S_{j}(T)\cdot g_{j}(\nu,T,p)
315: \end{equation}
316:
317: Here, $S_{j}(T)$ is the temperature-dependent line strength of a
318: particular spectral line $j$ and $g_{j}(\nu,T,p)$ the temperature-
319: and pressure-dependent line shape function of the same line.
320:
321: For the scattering coefficient, an analogous equation is valid:
322:
323: \begin{equation}\label{kappascattdefinition}
324: s_{\nu}=\sum_i s_{\nu,i}=\sum_i \sigma_{y,i}(\nu) \cdot N_i
325: \end{equation}
326:
327: Here, $\sigma_{y,i}(\nu)$ is a scattering cross section of type $y$.
328: In the present model, Rayleigh scattering is considered.
329:
330: The solution of eq. (\ref{opttiefe}), i.e. the calculation of
331: optical depths, is one of the key elements in radiative-convective
332: models. This solution, of course, depends on the accurate
333: calculation of the absorption cross sections.
334:
335: From eq. (\ref{endgueltig}), the necessary fluxes for eq.
336: (\ref{fluesse}) (i.e., the thermal and the solar flux) are obtained
337: by an angular integration of the (monochromatic) intensity:
338:
339: \begin{equation}
340: F_{\nu}=\int_{\Omega}^{}d\omega \mu I_{\nu} \label{angular}
341: \end{equation}
342:
343: and a frequency integration of the (monochromatic) flux:
344:
345: \begin{equation}
346: F_{\nu_1\rightarrow\nu_2}=\int_{\nu_1}^{{\nu_2}}F_{\nu}d\nu
347: \label{frequ}
348: \end{equation}
349:
350: Each one of these integrations is performed independently for the
351: two components of the total flux.
352:
353: \subsubsection{Solar radiation}
354:
355: The solar radiation module which calculates $F_{\rm{solar}}(z)$ for
356: eq. (\ref{fluesse}) has already been used by, e.g.,
357: \citet{pavlov2000,Seg2003} or \citet{Grenf2007a} and is based on
358: \citet{kasting1984} and \citet{kasting1988}. The module considers a
359: spectral range from 0.26 to 4.5 $\mu$m in 38 intervals. It evaluates
360: the solar incident radiation at a fixed daytime average solar zenith
361: angle of 60$^\circ$. Contributions to the optical depth come from
362: gaseous absorption by water and carbon dioxide (i.e., $\kappa_{\nu}$
363: in eq. (\ref{opttiefe})) and from Rayleigh scattering by carbon
364: dioxide, molecular nitrogen and molecular oxygen (i.e., $s_{\nu}$ in
365: eq. (\ref{opttiefe})).
366: %\newline
367:
368: Absorption cross sections $\sigma_{\rm{abs}}$ for the solar code
369: were obtained from the HITRAN 1992 database \citep{rothman1992}.
370: Rayleigh scattering cross sections $\sigma _{ray}$ are parameterized
371: following \citet{vardavas1984}. The frequency integration (see also
372: eq. (\ref{frequ})) of the RTE for $F_{solar}$ in each of the 38
373: spectral intervals is parameterized by a four-term correlated-k
374: exponential sum (e.g., \citealp{wiscombe1977}). The following
375: angular integration (eq. (\ref{angular})) is performed by using a
376: quadrature $\delta$-2-stream approximation code based on
377: \citet{Toon1989}. The resulting fluxes from each spectral interval
378: are added up to yield the total solar flux $F_{\rm{solar}}(z)$ at an
379: atmospheric level $z$. This flux is further multiplied by a factor
380: of 0.5 to account for diurnal variation.
381:
382: \subsubsection{Thermal molecular absorption}
383:
384: \label{MRAC}
385:
386: The thermal (planetary) radiation module for $F_{\rm{thermal}}(z)$
387: in eq. (\ref{fluesse}) considers a spectral range from 1 to 500
388: $\mu$m in 25 intervals. Our new thermal module is called MRAC
389: (Modified RRTM for Application in CO$_2$-dominated Atmospheres) and
390: is based on the radiation scheme RRTM (Rapid Radiative Transfer
391: Model). RRTM was developed by \citet{Mlawer1997} and has been used
392: by numerous other modeling studies (e.g.,
393: \citealp{Seg2003,Seg2005,Grenf2007a,Grenf2007b}). The need for a new
394: radiation model comes from the fact that RRTM was specifically
395: designed for conditions of modern Earth, i.e. it is not adaptable to
396: studies of atmospheres which greatly differ from modern atmospheric
397: conditions (in terms of atmospheric composition, temperature
398: structure, pressure, etc.). MRAC is easily adaptable to varying
399: conditions, as is described below.
400: %\newline
401:
402: MRAC uses the correlated-$k$ approach (e.g.,
403: \citealp{goody1989,lacis1991,Cola2003}) for the frequency
404: integration of the RTE in the thermal range, as does RRTM. The
405: planetary surface (bottom layer of the model atmosphere) and the
406: atmospheric layers are taken as blackbody emitters, according to
407: their respective temperatures. The thermal surface emissivity is set
408: to unity. The absorber species considered in MRAC are water, carbon
409: dioxide and carbon monoxide. The angular integration (eq.
410: (\ref{angular})) is performed using the diffusivity approximation
411: (as in \citealp{Mlawer1997}).
412:
413:
414: \textit{The correlated-k method}
415:
416:
417: Basically, this method transfers the frequency integration (FI) of
418: the RTE from frequency space $\nu$ to a probability space $g$. For
419: the absorption cross section $\sigma_{\rm{abs}}$ of eq.
420: (\ref{defabs}) in an interval $[\nu_1,\nu_2]$, a probabilistic
421: density distribution $f(\sigma_{\rm{abs}})$ (i.e., probability of
422: occurence for a particular value of $\sigma_{\rm{abs}}$) with the
423: following normalization condition can be defined:
424:
425: \begin{equation}
426: \int_{0}^{{\infty}}f(\sigma_{\rm{abs}})d\sigma_{\rm{abs}}=1
427: \label{fk}
428: \end{equation}
429:
430: To follow conventional nomenclature in the literature regarding
431: $k$-distributions, $\sigma_{\rm{abs}}$ is hereafter referred to as
432: $k$. The function $f(\sigma_{\rm{abs}})=f(k)$ is called the
433: $k$-distribution. From the $k$-distribution, a cumulative
434: $k$-distribution $g(k)$ can be defined by
435:
436: \begin{equation}
437: g(k)=\int_{0}^{{k}}f(k')dk' \label{gk}
438: \end{equation}
439:
440: The cumulative $k$-distribution is a strictly monotonic function and
441: may thus be inverted from $g(k)$ to yield $k(g)$. This mapping of
442: the frequency information ($k(\nu)$) into a single probability
443: variable ($k(g)$) can be done because it is irrelevant at which
444: position of the spectral interval a particular value of the
445: absorption cross section $k$ occurs. Performing a variable
446: substitution in eq. (\ref{frequ}) then leads to the following
447: equation:
448:
449: \begin{equation}
450: F_{\nu_1\rightarrow\nu_2}=\int_{\nu_1}^{{\nu_2}}F(\nu)d\nu=\int_{0}^{{1}}F(g)dg
451: \label{corrk}
452: \end{equation}
453:
454: The goal of the cumulative-$k$ approach is to reduce the number of
455: radiative transfer operations drastically while keeping the accuracy
456: of line-by-line models. This can be achieved with very few numbers
457: of points in $g$ space (see \citealp{goody1989,west1990}). The $g$
458: integration in eq. (\ref{corrk}) is performed in MRAC by Gaussian
459: quadrature using 16 intervals in $g$ space (the same as used in
460: RRTM, \citealp{Mlawer1997}).
461: %\newline
462:
463: The extension of this exact method from homogeneous to inhomogeneous
464: atmospheres is called the correlated-$k$ method (e.g.,
465: \citealp{Mlawer1997}). Each $g$ interval is treated as if it were a
466: monochromatic frequency interval, i.e. the method uses the same
467: subset of $g$ space for all layers throughout the atmosphere. This
468: implies a full frequency correlation of a specific subset of $g$
469: space for all atmospheric layers. There are conditions under which
470: this approach is exact \citep{goodyyung1989}, but usually these do
471: not hold. However, the numerical error of the correlated-$k$ method
472: is generally small \citep{Mlawer1997}.
473: %\newline
474:
475: \textit{Creating the new radiation scheme}
476: %\newline
477:
478: MRAC was originally designed to simulate atmospheres of a wide range
479: of possible terrestrial planets other than modern Earth, as stated
480: above. Therefore, a new temperature-pressure ($T$-$p$) grid to
481: incorporate the aforementioned ($T$,$p$)-dependence of the
482: absorption cross sections (see eq. \ref{defabs}) has been
483: introduced. In RRTM, $k$ values are tabulated for every spectral
484: band and every point in $g$ space for 59 pressure levels and the
485: associated Mid-Latitude-Summer (MLS) standard Earth temperature
486: values as well as temperature values T$_{\rm{MLS}}\pm$ 15 K and
487: T$_{\rm{MLS}}\pm$ 30 K, as described in \citet{Mlawer1997}. The
488: $T$-$p$ grid of RRTM is thus more or less fixed to modern Earth
489: conditions. For MRAC, we used 8 pressure levels, ranging
490: equidistantly in log $p$ from 10$^{-5}$ to 100 bar and 9 temperature
491: points, 6 in 50 K steps from 150 K to 400 K and three additional
492: points for 500, 600, 700 K respectively. Tests showed that this grid
493: allows for an interpolation accuracy of usually better than 2-3\%.
494: Furthermore, the number of $T$-$p$ points is consistent with
495: previous modeling studies \citep{kasting1993,Cola2003}.
496: \newline
497: Figure \ref{range} shows the range of the tabulated $k$-values for
498: both RRTM and MRAC, i.e. the interpolation regime of the two
499: radiative schemes. It demonstrates that RRTM can only be applied to
500: a much narrower range of atmospheric conditions in comparison to
501: MRAC.
502: %\newline
503:
504: %%FIGURE \ref{range}
505:
506: The necessary re-calculation of the $k$-distributions for each of
507: the gases included in MRAC (water, carbon dioxide, carbon monoxide)
508: proceeded in three steps:
509: %\newline
510:
511: First, the absorption cross sections for every species were
512: calculated for each $T$-$p$ grid point in each of the spectral
513: intervals where the respective species is active. The line shape
514: cut-off was set to 10 cm$^{-1}$ from the frequency $\nu$, i.e. the
515: sum over $j$ in eq. (\ref{defabs}) contains contributions from all
516: lines within $\pm$ 10 cm$^{-1}$. For the line shape
517: $g_{j}(\nu,T,p)$, a Voigt profile was assumed. The required line
518: parameters were taken from the HiTemp 1995 database
519: \citep{rothman1995}. The foreign broadening parameters in HiTemp are
520: given for air, i.e. an oxygen-nitrogen mixture as a background
521: atmosphere. However, as reported by several authors (e.g.,
522: \citealp{brown2005,toth2000}), the foreign broadening parameters
523: vary by significant amounts when different broadening gases are
524: considered. Thus, for each type of background atmosphere, a new set
525: of $k$-distributions must be generated (e.g., low-oxygen
526: atmospheres, CO$_2$-dominated atmospheres, intermediate
527: N$_2$-O$_2$-atmospheres). These line parameters were then used as
528: input to a line-by-line radiative transfer model called Mirart
529: \citep{schreier2003}. Mirart produced the actual absorption cross
530: sections with a spectral resolution of 10$^6$ equidistant points per
531: spectral interval.
532: %\newline
533:
534: Second, the $k$-distributions $f(k)$ were calculated from the
535: absorption cross sections. From the $k$ distribution $f(k)$, the
536: cumulative $k$-distribution $g(k)$ was then obtained.
537: %\newline
538:
539: Third, representative $k$ values were calculated for each of the $g$
540: subintervals. In this step, our algorithm followed the approach of
541: \citet{Mlawer1997}, i.e. for each of the 16 Gaussian subintervals in
542: $g$ space, an arithmetic mean absorption cross section was
543: calculated.
544: %\newline
545:
546: MRAC also implements a so-called binary species parameter $\eta$ for
547: transmittance calculations. This is employed in intervals having two
548: important absorber species (for more details, see
549: \citealp{Mlawer1997} or \citealp{Cola2003}):
550:
551: \begin{equation}\label{bsp}
552: \eta=\frac{C_1}{C_1+r\cdot C_2}
553: \end{equation}
554:
555: Here, $C_{1,2}$ are the concentrations of the two gases (in the case
556: of MRAC, water and carbon dioxide) and $r$ is some specified
557: reference ratio (mean modern Earth tropospheric values). Carbon
558: monoxide, although present in the model in six spectral intervals,
559: is not considered to be part of the binary species parameter. This
560: is partly due to the expected low concentrations in the simulated
561: atmospheres (typical theoretical and measured values for early Earth
562: and Mars are below 10$^{-4}$ volume mixing ratio), partly due to the
563: expected low temperatures (below 300 K), which means that the strong
564: CO fundamentals around 4 $\mu$m are completely outside the relevant
565: Planck emission windows. Consequently, carbon monoxide has only a
566: reduced impact on the radiation budget, compared to water and carbon
567: dioxide. $k$-distributions are calculated in MRAC for 5 different
568: values of $\eta$, ranging equidistantly from 0 (CO$_2$ only) to 1
569: (H$_2$O only).
570: %\newline
571:
572: Another improvement in MRAC, compared to RRTM, is the treatment of
573: the Planck function in each band. The fraction of thermal radiance
574: associated with a subset in $g$ space is calculated from eq. (11) in
575: \citet{Mlawer1997}:
576:
577: \begin{equation}\label{rrtmfg}
578: f_g=\frac{B_gw_g}{\overline{B}_g}
579: \end{equation}
580:
581: Here, $B_g$ is the average Planck function of the frequencies in the
582: subset of $g$ space, $w_g$ the Gaussian weight of the $g$ interval
583: and $\overline{B}_g$ the average Planck function of the whole
584: spectral band (i.e., the whole $g$ space). As temperature, pressure
585: and species concentrations vary, the different $g$ subsets will
586: correspond to different frequencies, and as such the value of $B_g$,
587: thus $f_g$, will vary. This has been taken into account while
588: constructing the $k$-distributions.
589:
590: \citet{Mlawer1997} tabulated values of $f_g$ for every value of the
591: binary species parameter and for two atmospheric reference levels,
592: one each in the troposphere and the stratosphere. In MRAC, values of
593: $f_g$ were tabulated for three temperatures and two pressure levels
594: as well as for the values of the binary species parameter. The most
595: important factor for $f_g$ is the binary species parameter $\eta$,
596: whereas the variation with temperature and pressure is rather small,
597: although not negligible. Therefore these 6 $T$-$p$ points are
598: regarded as to be sufficient.
599: %\newline
600:
601: A further difference between RRTM and MRAC is the distinction
602: between troposphere and stratosphere. In the troposphere, RRTM
603: changes major and minor absorbers in some of the spectral intervals
604: (see table 1 in \citealp{Mlawer1997}). In some spectral bands, no
605: absorption is considered in the stratosphere, in others, the number
606: of key species is reduced. This is done because on Earth, the
607: chemical and physical regimes are quite different in the troposphere
608: compared to those in the stratosphere. However, as this is mostly
609: due to Earth-specific conditions (e.g., the cold trap, tropopause
610: and temperature inversion all approximately occur at the same
611: altitude), this distinction was not incorporated into MRAC.
612: %\newline
613:
614: \subsubsection{Thermal continuum absorption}
615:
616: % add reference to kasting 1984 water cont PROOF
617:
618: Based on approximation formulations used by \citet{kasting1984},
619: \citet{kasting1984water} and \citet{Cola2003}, additional CO$_2$ and
620: H$_2$O continuum absorption in the thermal region is considered. In
621: contrast, the RRTM scheme only considers water continuum absorption
622: \citep{Mlawer1997}.
623:
624: Equation (\ref{co2contkasting}) shows the approximation for the
625: optical depth $ \tau_{\rm{cont},\rm{CO_2}}$ due to CO$_2$ continuum
626: absorption. The corresponding parameters are taken from
627: \citet{kasting1984}.
628:
629: \begin{equation}\label{co2contkasting}
630: \tau_{\rm{cont},\rm{CO_2}}=C_iW\cdot p_E\left(\frac{T_0}{T}\right)^{t_i}
631: \end{equation}
632:
633: In this equation, $C_i$ a frequency-dependent adjustment to the path
634: length, $W$ the column amount of CO$_2$, $p_E=(1+0.3\cdot
635: C_{\rm{CO_2}})\cdot p$ ($p$ layer pressure, $C_{\rm{CO_2}}$
636: concentration) an effective CO$_2$ broadening pressure and $T_0$=300
637: K is a reference temperature.
638:
639:
640: The optical depth $\tau_{\rm{cont},\rm{H_2O}}$ due to water
641: continuum absorption in the window region (8-12 $\mu$m) is
642: calculated from the equation \citep{kasting1984water}
643:
644: \begin{equation}\label{waterkont}
645: \tau_{\rm{cont},\rm{H_2O}}=h_n\cdot p \cdot \frac{W_w^2}{W_t}
646: \end{equation}
647:
648: where $h_n=h_t\cdot h_{\nu}$ incorporates the frequency and
649: temperature dependence, $p$ is the pressure, $W_t$ the total and
650: $W_w$ the water column of the layer. $h_{\nu}$ is evaluated at the
651: high frequency interval boundary, as in \citet{kasting1984water}. We
652: use the following approximations for $h_t$ and $h_{\nu}$, based on
653: \citet{kasting1984water} and \citet{Cola2003}:
654:
655: \begin{equation}\label{approxtempwater}
656: % add factor of 1.800, omitted earlier PROOF
657: h_t=e^{1800\cdot(\frac{1}{T}-\frac{1}{296})}
658: \end{equation}
659: \begin{equation}\label{approxfreqwater}
660: h_{\nu}=1.25\cdot 10^{-22}+1.67\cdot 10^{-19}\cdot e^{-2.62\cdot 10^{-13}\cdot \nu}
661: \end{equation}
662:
663: Both the water and the carbon dioxide continuum absorption are
664: considered to be approximately monochromatic over a specific
665: spectral interval, hence their contribution to the overall
666: absorption coefficient (see eq. (\ref{opttiefe})) is added as a
667: constant term.
668:
669: \subsubsection{Convective adjustment}
670:
671: \label{convadjust}
672:
673: Convective adjustment to the lapse rate is performed whenever the
674: calculated radiative lapse rate $\nabla_{\rm{rad}}T$ exceeds the
675: adiabatic value $\nabla_{\rm{ad}}T$ (Schwarzschild criterion):
676:
677: \begin{equation}
678: \nabla_{\rm{rad}}T>\nabla_{\rm{ad}}T\label{schwarz}
679: \end{equation}
680:
681: The adiabatic lapse rate is calculated as a standard dry adiabat in
682: the stratosphere. In the troposphere, a wet H$_2$O adiabatic lapse
683: rate is assumed. Below 273 K, the Clausius-Clapeyron-equation
684: $\frac{d \ln(p_v)}{d \ln(T)}=\frac{m\cdot L}{R\cdot T}$ ($R$
685: universal gas constant, $m$ mass, $L$ latent heat release per mass)
686: for the saturation vapor pressure curve $p_v$ is applied. Between
687: 273 and 647 K, a formulation by \citet{ingersoll1969} is taken.
688:
689: \subsection{Atmospheric water profile}
690:
691: \label{watcalcu}
692:
693:
694: In every time step, the water vapor profile is re-calculated
695: according to the new temperature profile.
696:
697: In the troposphere, water vapor concentrations $C_{\rm{H_2O}}$ are
698: calculated from a fixed relative humidity distribution $RH$:
699:
700: \begin{equation}\label{relhumwater}
701: C_{\rm{H_2O}}(T,z)=\frac{p_{\rm{sat}}(T(z))}{p(z)}\cdot RH(z)
702: \end{equation}
703:
704: where $p_{\rm{sat}}$ is the saturation vapor pressure of water at
705: the given temperature $T$ and $p$ the atmospheric pressure at level
706: $z$. The default relative humidity profile $RH$ follows the approach
707: of \citet{manabewetherald1967}, with a relative humidity $R_s$ of
708: 80\% at the surface.
709:
710: \begin{equation}\label{manabewetherald}
711: RH(z)=R_s \cdot \frac{\frac{p(z)}{p_{\rm{surface}}}-0.02}{0.98}
712: \end{equation}
713:
714: Above the cold trap, water vapor is treated as a non-condensable
715: gas, and its concentration is fixed at the cold trap value.
716: %\newline
717:
718: \subsection{Boundary conditions, initial values and parameters}
719:
720: \label{summarymodelinput}
721:
722:
723: Since eq. (\ref{timestep}) is a first order differential equation
724: for the temperature, a starting temperature profile must be
725: provided. In addition, a boundary condition for the radiative flux
726: must be specified. To obtain unique equilibrium solutions,
727: parameters must also be provided for the model. These include
728: pressure parameters for the planetary top-of-atmosphere (TOA)
729: pressure $p_0$, gas concentrations, surface albedo or solar zenith
730: angle, for example.
731:
732: Table \ref{paraminput} summarizes the boundary conditions, initial
733: values and parameters.
734:
735: %TABLE \ref{paraminput}
736:
737:
738:
739: \section{Validation of the new radiation scheme}
740:
741: \label{validmracsubsection}
742:
743: \subsection{General remarks}
744:
745: MRAC has been tested in two different ways.
746:
747: \begin{itemize}
748: \item Case 1: $k$-distributions
749: \newline
750: The calculated $k$-distributions, i.e. the model input data, have
751: been validated against published values to show that the algorithm
752: creating the $k$-distributions works correctly.
753:
754: \item Case 2: Earth atmosphere temperature profiles
755: \newline
756: Temperature profiles of an Earth-like test atmospheres (composition:
757: N$_2$ 0.77, O$_2$ 0.21, Ar 0.01, CO$_2$ 3.55 $\cdot$ 10$^{-4}$)
758: calculated with MRAC and RRTM have been compared. This was done
759: since RRTM has been extensively validated both against line-by-line
760: codes and atmospheric measurements \citep{Mlawer1997} under modern
761: Earth conditions.
762:
763: Our test atmosphere has a composition close to the present day
764: atmosphere. However, it lacks radiative trace gases such as nitrous
765: oxide, ozone and methane, as these gases cannot be handled by MRAC
766: yet (see above). Note that due to the lack of ozone in our test
767: atmosphere, we do not expect a large stratospheric temperature
768: maximum as is observed in the present Earth atmosphere because this
769: maximum is almost entirely due to the absorption of solar radiation
770: by ozone.
771:
772: We additionally performed test runs on a second test atmosphere (not
773: shown) which differs from the first one by its CO$_2$ content
774: (100-fold increase). This 100-fold increase in CO$_2$ represents the
775: current limit for the RRTM scheme (\citealp{Seg2003}, Eli Mlawer,
776: priv. comm.).
777:
778: \end{itemize}
779:
780:
781: These two validation approaches are discussed below.
782:
783:
784: \subsection{$k$-distributions}
785:
786:
787: Figures \ref{kdislacis} and \ref{kdismlawer} compare our calculated
788: $k$-distributions (dotted lines) with previously published values
789: (plain lines) for different water and carbon dioxide bands.
790: Published values were taken from \citet{lacis1991} (H$_2$O, Fig.
791: \ref{kdislacis}) and \citet{Mlawer1997} (CO$_2$, Fig.
792: \ref{kdismlawer}).
793: %FIGURE \ref{kdislacis}
794: %FIGURE \ref{kdismlawer}
795: Figures \ref{kdislacis} and \ref{kdismlawer} indicate quite good
796: agreement with the published values.
797:
798:
799: \subsection{Earth temperature profiles}
800:
801:
802: The calculated temperature profiles for the test atmosphere are
803: shown in Fig. \ref{MRACvalid}.
804:
805: %FIGURE \ref{MRACvalid}
806:
807:
808: Figure \ref{MRACvalid} implies some differences in the middle to
809: upper stratosphere (2-6 K) and small deviations ($\ll1 K$) below
810: about 20 km. For the test atmosphere with a 100-fold increase in
811: CO$_2$ (not shown), the stratospheric differences are even larger
812: (up to 10 K). We interpret these differences in the temperature
813: profiles as follows:
814:
815: Firstly, as stated above in section \ref{MRAC}, MRAC does not
816: differentiate between troposphere and stratosphere, as is the case
817: for RRTM. That means, spectral bands where H$_2$O or CO$_2$ absorb
818: only weak are not considered for optical depth calculations in the
819: stratosphere by RRTM (e.g., bands 6, 12-13 and 15-16). In contrast,
820: MRAC incorporates the contribution of CO$_2$ and H$_2$O to the
821: optical depth in these spectral bands. However, this contribution is
822: usually rather small.
823: %\newline
824:
825: Secondly, and more importantly, the differences between the
826: stratospheric temperature profiles occur where RRTM has to use a
827: temperature extrapolation for the absorption cross sections beyond
828: the limits of its tabulated values.
829:
830: Fig. \ref{limitstable}, shows the temperature profile for our test
831: atmosphere Earth 1. Also shown is the validity range of RRTM as
832: already indicated in Fig. \ref{range}. This represents the lower
833: temperature limit for the tabulated absorption cross sections in
834: RRTM, as stated above.
835:
836: As can be seen from Fig. \ref{limitstable}, the calculated
837: temperature values for the first test atmosphere are below the lower
838: RRTM validity limit. Therefore, RRTM uses linear extrapolation to
839: calculate the absorption cross sections. This introduces a large
840: extrapolation error. On average, the calculated absorption cross
841: sections are a factor of 2-5 too low, depending on the spectral
842: band. Sometimes, the extrapolation performed by RRTM even yields
843: negative absorption cross sections. The interpolation errors in
844: MRAC, on the contrary, reach only 1-2\% on average.
845:
846: %FIGURE \ref{limitstable}
847:
848:
849: Figure \ref{fluxdiff} shows the radiative fluxes and heating and
850: cooling rates calculated by RRTM and MRAC in the test atmosphere
851: Earth 1. Solar fluxes and heating rates differ by much less than 1
852: \%. The thermal down-welling fluxes calculated by RRTM and MRAC show
853: large differences in the stratosphere below pressures of around
854: 10$^{-2}$-10$^{-3}$ bar, reaching up to a factor of 5 in the upper
855: stratosphere where pressures are below 10$^{-4}$ bar. However, these
856: differences are well below 1 W m$^{-2}$, so are not discernible on
857: the scale in Figure \ref{fluxdiff}. The up-welling fluxes differ by
858: only about 10 \% in the stratosphere, since they are dominated by
859: the tropospheric component. Hence, the calculated resulting cooling
860: rates differ only by small amounts of 0.1-0.4 K day$^{-1}$ and
861: usually lie within 5-10 \%, especially in the upper stratosphere.
862:
863: %FIGURE \ref{fluxdiff}
864:
865: In order to assess the sensitivity of the model to errors in the
866: absorption cross sections (hence, in optical depth and thermal
867: fluxes), we artificially increased the optical depth in RRTM in the
868: most important stratospheric band, the CO$_2$ 15$\mu$m fundamental
869: by factors of 2, 5, 10 and 20, respectively. Fig. \ref{vartau}
870: quantifies the effect of these sensitivity runs on the temperature
871: profile. Fig. \ref{vartau} a) shows the total optical depth
872: calculated by RRTM and MRAC in the validation runs. Clearly, in the
873: stratosphere, RRTM under-estimates the optical depth, as already
874: discussed above.
875:
876: %FIGURE \ref{vartau}
877:
878: Fig. \ref{vartau} b) shows the temperature profiles for the two
879: Earth validation runs with MRAC and RRTM, as well as for a run with
880: RRTM, but increased optical depth by a factor of 2. This factor of 2
881: is representative of the error in the absorption cross section
882: calculations in the 15 $\mu$m band due to the required extrapolation
883: in the $T$-$p$-range, as shown in Figure \ref{range}. It can be seen
884: that by increasing the optical depth in RRTM artificially, the
885: temperature profile nearly reproduces the MRAC temperature profile.
886:
887:
888:
889: These results show that conditions which differ too much from the
890: Earth's standard atmosphere seem to pose problems for RRTM. This
891: limitation was already noted in some of the previous studies
892: performed with RRTM. Clearly, due to the use of an expanded
893: temperature range, MRAC performs better than RRTM in these
894: atmospheres.
895: %\newline
896:
897:
898: \section{About the runs}
899:
900: \label{about}
901:
902: \subsection{Absorption cross sections}
903:
904: The absorption cross sections used in the runs performed for this
905: work (summarized in Tables \ref{runssummary} and \ref{sensruns})
906: were calculated assuming a N$_2$-CO$_2$-background atmosphere,
907: consisting of 95\% molecular nitrogen and 5\% carbon dioxide.
908: According to \citet{kasting1986} and \citet{toth2000}, the foreign
909: broadening coefficient for water is enhanced by a factor of 2 with
910: respect to air for CO$_2$ as a broadening gas and by a factor of 1.2
911: for N$_2$ as a broadening gas. Similarly, for carbon dioxide the
912: foreign broadening coefficient was enhanced by a factor of 1.3 when
913: N$_2$ was the broadening gas \citep{kasting1986}. Accounting for the
914: appropriate mixing ratios then yields an effective enhancement
915: factor by which the foreign broadening parameter from HiTemp was
916: multiplied before use in the cross section calculations described
917: above.
918:
919: We compared the calculated cross sections of the assumed
920: N$_2$-CO$_2$-atmosphere (95\% nitrogen and 5\% carbon dioxide) with
921: cross sections for some major spectral bands corresponding to
922: different background atmospheres, i.e. different CO$_2$
923: concentrations. These cross sections were obtained with the
924: line-by-line radiative transfer model Mirart \citep{schreier2003}.
925: The cross sections from the 95\%-N$_2$-5\%-CO$_2$-atmosphere agree
926: within 5 \% for most of the cases studied in this work, although for
927: the runs with the lowest CO$_2$ concentrations, the agreement
928: decreased to about 10 \%.
929: \newline
930:
931:
932: \subsection{ Model runs}
933:
934: There were a total number of 12 nominal runs performed as shown in
935: Table \ref{runssummary}.
936:
937: We assumed a constant background pressure of 0.77 bar N$_2$ with
938: variable amounts of CO$_2$. Water vapour contents of the atmosphere
939: were calculated as described in section \ref{watcalcu}. No other
940: gases were present in the atmosphere, as in \citet{kasting1987}. For
941: several values of the solar constant ($S$=0.7, 0.75, 0.8, 0.85, 0.9
942: and 0.95 present-day value), we increased the CO$_2$ partial
943: pressure until converged surface temperatures reached 273 K (runs
944: 1-6) and 288 K (runs 7-12), respectively. This procedure is similar
945: to what was done by \citet{kasting1987}. The solar constant values
946: were chosen as to loosely correspond to important events throughout
947: the Earth's history, such as the beginning of the main sequence life
948: time of the Sun ($S$=0.70, 4.6 Gy ago), the end of the late heavy
949: bombardment ($S$=0.75, 3.8 Gy ago), the rise of oxygenesis by
950: cyanobacteria ($S$=0.80, 2.9 Gy ago), the first and the second
951: oxidation event ($S$=0.85 and $S$=0.90, 2 and 1.3 Gy ago
952: respectively) and the Cambrian explosion ($S$=0.95, 0.6 Gy ago).
953: Assigning geological ages to the solar constants used in the model
954: (see Table \ref{runssummary}) is essential when comparing the
955: calculated model CO$_2$ concentrations to the available data and
956: constraints on carbon dioxide. In this work, we chose two
957: approximations of the solar luminosity $S(t)$ with time. The first
958: one is from \citet{caldeira1992}, the second approximation is from
959: \citet{gough1981}. The difference of these two formulations is most
960: pronounced for earlier time periods before 1-2 Gy ago, although it
961: is rarely larger than a few percent.
962:
963: The total surface pressure is always calculated from the relation
964: $p_{\rm{surf}}=p_{\rm{CO_2}}+p_{\rm{N_2}}+p_{\rm{H_2O}}$. Since we
965: calculated surface temperatures of 273 K and 288 K, the amount of
966: surface pressure which arose from the evaporation of water is about
967: 6 mb and 17 mb, respectively. Note that the total surface pressure
968: in our model runs decreases with time, as we assume less CO$_2$
969: partial pressure in our model atmospheres. There is no geological
970: data available on the total surface pressure throughout time,
971: however, our approach of constant N$_2$ partial pressure is
972: consistent with assumptions made in previous studies (e.g.,
973: \citealp{kasting1987}).
974:
975: %%TABLE \ref{runssummary}
976:
977:
978:
979:
980:
981: \subsection{Parameter variations of surface albedo and relative humidity profile}
982:
983: Previous works regarding the "faint young Sun problem" have
984: concentrated on the increased greenhouse effect, as stated above in
985: the Introduction. We also studied two important parameters affecting
986: the surface temperature in our model, namely the surface albedo and
987: the relative humidity (RH) distribution.
988:
989: The importance of cloud coverage for the surface temperature, e.g.
990: on early Mars, has been studied by \citet{mischna2000}. The effect
991: in their model was quite large, yielding surface temperatures from
992: 220 K to 290 K, depending on cloud optical depth and cloud location.
993: In our model, the surface albedo simulates the presence of clouds,
994: as stated above in the description of our model.
995:
996: Water vapour is a very effective greenhouse gas. In contrast to
997: CO$_2$, its content in the troposphere (where 99 \% of the column
998: resides) is controlled by the hydrological cycle (ocean reservoir,
999: evaporation, subsequent condensation and precipitation) which is
1000: very sensitive to temperature. A critical parameter in climate
1001: models is the relative humidity parametrization, which clearly has a
1002: potentially large impact on the water content in the atmosphere and
1003: hence on the resulting greenhouse effect.
1004: %\newline
1005:
1006: In addition to the model runs described above, we therefore
1007: performed runs varying surface albedo and relative humidity
1008: profiles. These runs are based on runs 3 and 4 of Table
1009: \ref{runssummary}, i.e. solar constants of $S$=0.8 and $S$=0.85.
1010: %\newline
1011: Table \ref{sensruns} summarizes the parameter variations.
1012:
1013:
1014: %%TABLE \ref{sensruns}
1015:
1016: In the runs from Table \ref{runssummary}, the surface albedo is
1017: normally set to $A$=0.21 (see Table \ref{paraminput}). Now, we
1018: assume surface albedos of $A$ $\pm$10 \%, i.e. 0.19 and 0.23,
1019: respectively.
1020:
1021: The relative humidity profile used for the calculation of water
1022: vapour concentrations in Table \ref{runssummary} follows the
1023: approach of \citet{manabewetherald1967} (referred to as RH=MW, see
1024: section \ref{watcalcu}). Here, for the additional parameter studies,
1025: we used three different RH profiles. The first one assumed a
1026: saturated troposphere (RH=1). The second profile used RH=0, i.e.
1027: water was removed from the atmosphere. These two profiles represent
1028: the lower and upper limits of the atmospheric water content. The
1029: third RH profile is a more realistic RH profile. It is based on a
1030: temperature correction to the RH profile of
1031: \citet{manabewetherald1967} which was first proposed by
1032: \citet{cess1976} and later used by \citet{vardavas1985}.:
1033:
1034: \begin{equation}\label{relhumcess}
1035: R=R_{\rm{surface}}\cdot R_{\rm{mw}}^{1-0.03(T_S-288)}
1036: \end{equation}
1037:
1038: where $R_{\rm{surface}}\cdot R_{\rm{mw}}$ is the relative humidity
1039: distribution from eq. \ref{manabewetherald}. This profile is
1040: referred to as RH=C. Figure \ref{relhumdif} shows the difference
1041: between the two RH profiles from \citet{manabewetherald1967} and
1042: \citet{vardavas1985}. The total amount of water vapour is reduced by
1043: about 20 \% by using the temperature correction of
1044: \citet{vardavas1985} compared to \citet{manabewetherald1967}.
1045: %\newline
1046:
1047: %%FIGURE \ref{relhumdif}
1048:
1049:
1050:
1051:
1052: \section{Results and Discussion}
1053:
1054: \label{showresults}
1055:
1056: \subsection{Examples of thermal structure and water profiles}
1057:
1058:
1059: Figure \ref{tprofile} shows the temperature profiles for run 3
1060: (plain line) and run 4 (dotted line). Run 3 ($S=0.80$) considered 20
1061: mb CO$_2$ partial pressure, run 4 ($S=0.85$) roughly 3 mb (see Table
1062: \ref{runssummary}). These two runs were chosen because they
1063: represent interesting points in time, such as the advent of
1064: cyanobacteria (run 3) and the first oxidation event (run 4).
1065:
1066: %FIGURE \ref{tprofile}
1067:
1068: Clearly, in the troposphere the temperature differences for the two
1069: runs are rather small. In the lower to middle stratosphere (from
1070: around 10 to 25 km), run 4 (dashed line) shows lower temperatures
1071: than run 3 (plain line). In this region, absorption of solar
1072: radiation by CO$_2$ is the dominant heating process, eventually
1073: causing a temperature inversion from the middle stratosphere upwards
1074: (seen for both runs in Fig. \ref{tprofile} for higher altitudes
1075: above 25-30 km). In the upper stratosphere (above 25-30 km), where
1076: CO$_2$ radiative cooling via the 15 $\mu$m fundamental band sets in,
1077: run 3 shows lower temperatures than run 4. This reflects the lower
1078: flux and higher CO$_2$ concentrations in run 3 compared to run 4.
1079:
1080: However, as already noted by \citet{Seg2003}, the actual strength of
1081: the temperature inversion is not well determined due to less
1082: accurate transmittance calculations in the solar code and may well
1083: turn out to be a mere numerical problem rather than reflecting
1084: physical processes.
1085: %\newline
1086:
1087: Also indicated in Fig. \ref{tprofile} are the convective zones for
1088: both runs. The convective zone for the $S=0.8$ case only reaches up
1089: to about 3 km altitude, whereas in the $S=0.85$ case, it extends
1090: already to about 7.5 km, which is closer to the present-day,
1091: latitude-dependent value of 10-20 km. However, the lapse rate
1092: remains close to the adiabatic value for several kilometres after
1093: entering the regime of radiative equilibrium (above the dot-dashed
1094: lines in Fig. \ref{tprofile}).
1095: %\newline
1096:
1097: Figure \ref{wprofile} is similar to Fig. \ref{tprofile}, but it
1098: shows the resulting water profiles. There is considerably less water
1099: in the stratosphere for the $S=0.85$ case due to enhanced
1100: condensation. The cold trap position is indicated, as well as the
1101: tropopause position, i.e. the boundary between the convective and
1102: non-convective regime.
1103:
1104: %FIGURE \ref{wprofile}
1105:
1106: Figures \ref{tprofile} and \ref{wprofile} illustrate some
1107: interesting features regarding the thermal structure of the
1108: atmosphere. The cold trap, i.e. the highest point up to which water
1109: is allowed to condense out in the model, is no longer located near
1110: the tropopause, as is the case in the atmosphere of modern Earth. It
1111: is still associated with a temperature inversion, but as this
1112: inversion occurs at higher altitudes, cold trap and tropopause
1113: (i.e., the boundary between convective and non-convective layers)
1114: are located at noticeably different altitudes. It seems that the
1115: observation that tropopause, cold trap and temperature inversion
1116: occur at approximately the same height in the modern Earth's
1117: atmosphere is somewhat coincidental.
1118: %\newline
1119:
1120: \subsection{Climatic constraints on CO$_2$ partial pressures}
1121:
1122:
1123: The results of the 12 nominal runs (see Table 4) are summarized in
1124: Fig. \ref{minin2data}. The lower plain line indicates the CO$_2$
1125: partial pressures corresponding to calculated surface temperatures
1126: of 273 K (runs 1-6), the upper plain line shows the partial
1127: pressures for surface temperatures of 288 K (runs 7-12). For
1128: comparison, the dashed line shows the partial pressures as
1129: calculated by \citet{kasting1987}. Fig. \ref{minin2data} shows that
1130: our new radiative scheme requires considerably less CO$_2$ (about a
1131: factor of 2-15) to achieve an ice-free surface than the study of
1132: \citet{kasting1987}.
1133: % add this sentence PROOF
1134:
1135: %\newline
1136:
1137: %FIGURE \ref{minin2data}
1138: Note that the radiative transfer scheme used by \citet{kasting1987}
1139: was applicable to this type of atmospheres, as is the one we used.
1140:
1141:
1142:
1143: Also indicated in Fig. \ref{minin2data} are the upper limits of
1144: CO$_2$ partial pressures as inferred from sedimentary data.
1145:
1146: Several attempts have been made to constrain the CO$_2$ content of
1147: the early Earth atmosphere. For example, among others,
1148: \citet{rye1995}, \citet{hessler2004} and \citet{towe1985} have
1149: placed upper limits on CO$_2$ partial pressures during different
1150: periods of the mid-Archaean to early Proterozoic of approximately
1151: 10-100 PAL (Present Atmospheric Level), depending on temperature.
1152: These limits were based on experimental sediment data and on
1153: biological considerations. \citet{rye1995} and \citet{hessler2004}
1154: argued for upper limits on CO$_2$ partial pressures based on the
1155: observations that in ancient Archaean rocks, siderites are missing.
1156: \citet{towe1985} stated atmospheric upper levels of CO$_2$ based on
1157: the observation that anaerobic photosynthesis and nitrogen fixation
1158: (which are both believed to be evolutionary ancient) are
1159: incompatible with high CO$_2$ concentrations.
1160: %\newline
1161:
1162: Previous climate studies of the early Earth's atmosphere calculated
1163: very high CO$_2$ partial pressures in order to raise the surface
1164: temperature above 273 K. The model CO$_2$ values were generally
1165: above 50 mb well into Proterozoic age (e.g., Kasting 1987, see also
1166: Fig. \ref{minin2data}). However, the experimental data from
1167: sediments and biology (see above) were of the order of several mb,
1168: i.e. an order of magnitude lower. This contradiction is known as the
1169: "faint young Sun problem".
1170:
1171: Fig. \ref{minin2data} however shows that for the late Archaean
1172: (solar constant S=0.85), our new results are compatible with the
1173: paleosol records, hence the "faint young Sun problem" might be
1174: resolved for this time period.
1175: %\newline
1176:
1177: One should note that the absence of methane, ozone, ammonia or other
1178: greenhouse gases in our results implies that we calculated only a
1179: lower limit for surface temperatures. Photochemical models of the
1180: anoxic Archaean and low-oxygen Proterozoic atmospheres have shown
1181: that methane could build up to concentrations of the order of
1182: 10$^{-3}$ \citep{pavlov2003} which could also have contributed to
1183: the greenhouse effect. \citet{haqq2008} proposed higher hydrocarbons
1184: such as C$_2$H$_6$ in a methane-rich atmosphere as radiative gases.
1185: Those effects were not investigated here. Our studies imply,
1186: however, that much less to none additional greenhouse gases are
1187: required to warm the early Earth.
1188:
1189: \subsubsection{Effect of parameter variations of surface albedo and relative humidity}
1190:
1191: Table \ref{sensruns} summarizes the results of the parameter
1192: variations. Shown are the values of CO$_2$ partial pressure required
1193: to obtain the desired surface temperature of 273 K.
1194:
1195: Upon lowering the surface albedo value from 0.23 to 0.19, the
1196: partial pressure of CO$_2$ required to keep the surface at 273 K is
1197: reduced from 28.5 mb to 11.0 mb (at $S$=0.80) and from 5.5 mb to 1.5
1198: mb (at $S$=0.85). This is a lowering of the amount of CO$_2$ by
1199: factors of 2.6 and 3.7, respectively.
1200:
1201: When changing the RH profile from saturated (RH=1) to water-free
1202: conditions (RH=0), the necessary partial pressure of CO$_2$ has to
1203: be increased from 4.9 mb to 266.8 mb ($S$=0.80) and from 0.6 mb to
1204: 180.6 mb ($S$=0.85). However, like expected, the comparison between
1205: the more realistic RH profiles of \citet{manabewetherald1967} and
1206: \citet{cess1976} yields much smaller differences. When changing the
1207: RH profile from RH=MW to RH=C, CO$_2$ partial pressures must be
1208: increased from 19.1 mb to 27.3 mb ($S$=0.80) and from 2.9 mb to 4.9
1209: mb ($S$=0.85). These differences amount to 43 \% and 69 \%,
1210: respectively, which is much lower than the differences obtained by
1211: varying the surface albedo.
1212: %\newline
1213:
1214: %\subsubsection{Importance of additional parameters}
1215:
1216: These parameter variations demonstrated that the surface albedo,
1217: hence clouds, is a very important parameter in view of the surface
1218: temperature. The results obtained imply that the latter is probably
1219: more important than the RH profile, although the RH profile has to
1220: be incorporated more consistently to accurately constrain CO$_2$
1221: partial pressures for the "faint young Sun" problem. This calls for
1222: a more elaborated 1D model incorporating clouds and cloud formation
1223: as was done by, e.g., \citet{Cola2003} for early Mars. In the
1224: future, we plan to add clouds to our code.
1225: %\newline
1226:
1227: The results of the parameter variations, however, did not
1228: significantly change our main conclusion, namely that the CO$_2$
1229: values for the late Archaean calculated by our improved model are
1230: consistent with observational data. The CO$_2$ values obtained for
1231: the parameter variations range between 1.5-5.5 mb of partial
1232: pressure. This is still close to the values inferred from the
1233: paleosol record.
1234:
1235:
1236:
1237:
1238: \section{Summary}
1239:
1240: \label{conclusions}
1241:
1242: In this work, we addressed the "faint young Sun problem" of the
1243: ice-free early Earth. In order to do this, we applied a
1244: one-dimensional radiative-convective model to the atmosphere of the
1245: early Earth. Our model included updated absorption coefficients in
1246: the thermal radiative transfer scheme.
1247:
1248: The validations done for the new radiative transfer scheme have been
1249: described. The new scheme is found to perform significantly better
1250: than
1251: % change to refer only to RRTM PROOF
1252: a previous scheme under conditions deviating from the modern
1253: Earth atmosphere.
1254:
1255: We then studied the effect of enhanced carbon dioxide concentrations
1256: and parameter variations of the surface albedo and the relative
1257: humidity profiles on the surface temperature of the early Earth with
1258: the improved model.
1259:
1260: Our new model simulations suggest that the amount of CO$_2$ needed
1261: to keep the surface of the early Earth from freezing is
1262: significantly less (up to an order of magnitude) than previously
1263: thought (see Figure \ref{minin2data}).
1264:
1265: For the late Archaean and early Proterozoic period (around 2-2.5 Gy
1266: ago), the calculated amount of CO$_2$ (2.9 mb partial pressure)
1267: which is needed to obtain a surface temperature of 273 K is
1268: compatible with the amount inferred from geological data, contrary
1269: to previous studies (see Figure \ref{minin2data}). The apparent
1270: contradiction between model constraints on CO$_2$ and sediment data
1271: disappears for this time period.
1272:
1273: Upon varying model parameters such as the surface albedo and
1274: relative humidity profile, we found this conclusion to be robust.
1275: The calculated CO$_2$ partial pressures for the late Archaean
1276: (1.5-5.5 mb) are still consistent with the geological evidence.
1277:
1278: \section*{Acknowledgements}
1279:
1280: We are grateful to Jim Kasting and Eli Mlawer for useful discussion
1281: while creating the new radiation scheme. Furthermore, we are
1282: grateful to Viola Vogler for help in doing some of the plots in this
1283: paper.
1284:
1285: We thank the two anonymous referees for their constructive remarks
1286: which helped to improve and clarify this paper.
1287:
1288:
1289: \newpage
1290:
1291:
1292: % bibliography inclusion
1293: \bibliographystyle{elsart-harv}
1294: \bibliography{v2}
1295:
1296: \newpage
1297: \linespread{1.}
1298:
1299: \section*{Figure captions}
1300:
1301:
1302: Figure \ref{range}:
1303:
1304: Range of $T$-$p$ values used to obtain absorption cross sections in
1305: the two radiative schemes RRTM (light grey) and MRAC (dark grey).
1306: \newline
1307:
1308:
1309: Figure \ref{kdislacis}:
1310:
1311: Cumulative $k$ distribution for parts of the 6.3 $\mu$m H$_2$O
1312: fundamental band, T=240 K and p=10 mb: Comparison between
1313: \citet{lacis1991} (plain line) and
1314: our algorithm (dotted).
1315: \newline
1316:
1317: Figure \ref{kdismlawer}:
1318:
1319: Cumulative $k$ distribution for parts of the 15 $\mu$m CO$_2$
1320: fundamental band between
1321: 630 and 700 cm$^{-1}$, T=260 K and p=507 mb: Comparison between \citet{Mlawer1997} (plain line)
1322: and our algorithm (dotted).
1323: \newline
1324:
1325: Figure \ref{MRACvalid}:
1326:
1327: Comparison of validation temperature profiles (RRTM plain line, MRAC
1328: dotted line).
1329: \newline
1330:
1331: Figure \ref{limitstable}:
1332:
1333: Limits of the RRTM temperature grid. Atmospheric conditions as for
1334: the standard CO$_2$ case in fig. \ref{MRACvalid}, RRTM profile
1335: (plain), MRAC (dotted). The shaded area designates the validity
1336: range of RRTM as already indicated in fig. \ref{range}.
1337: \newline
1338:
1339:
1340:
1341:
1342: Figure \ref{fluxdiff}:
1343:
1344: Flux profiles (thermal and solar up- and down-welling fluxes) (left
1345: panel) and heating and cooling rates (right panel) calculated by the
1346: two radiative schemes in the test atmosphere.
1347: \newline
1348:
1349: Figure \ref{vartau}:
1350:
1351: Effect of variations in the optical depth on the temperature profile
1352: calculated by RRTM: Left panel, total optical depth, right panel,
1353: temperature profiles.
1354: \newline
1355:
1356:
1357:
1358: Figure \ref{relhumdif}:
1359:
1360: Differences of the two relative humidity profiles used for the
1361: parameter variations
1362: \newline
1363:
1364:
1365: Figure \ref{tprofile}:
1366:
1367: Temperature profiles for runs with solar constants of $S=0.8$ (run
1368: 3, plain line) and $S=0.85$ (run 4, dotted); the background pressure
1369: is 0.77 bar of N$_2$. The convective zones are indicated for both
1370: runs.
1371: \newline
1372:
1373: Figure \ref{wprofile}:
1374:
1375: Water profiles for runs with solar constants of $S=0.8$ (run 3,
1376: plain line) and $S=0.85$ (run 4, dashed); the background pressure is
1377: 0.77 bar of N$_2$. Cold trap positions and tropopause positions are
1378: indicated for both runs.
1379: \newline
1380:
1381:
1382: Figure \ref{minin2data}:
1383:
1384: Minimal values of CO$_2$ partial pressure required to obtain chosen
1385: surface temperatures of 273 K and 288 K at a fixed N$_2$ partial
1386: pressure for different solar constants: shown are the curves for the
1387: new model (plain lines, upper line 288 K, lower line 273 K) and the
1388: model of \citet{kasting1987} (dashed)). Symbols are included which
1389: represent the upper CO$_2$ limits derived from the sediment record
1390: ($\Box$: age conversion by \citet{caldeira1992}, $\triangle$: age
1391: conversion by \citet{gough1981}.
1392:
1393:
1394: \newpage
1395: \section*{Tables}
1396:
1397: \input{table1}
1398: \newpage
1399:
1400: \input{table2}
1401: \newpage
1402:
1403: \input{table3}
1404: \newpage
1405:
1406: \input{table4}
1407: \newpage
1408:
1409:
1410:
1411: \section*{Figures}
1412:
1413:
1414: \begin{figure}[H]
1415: % % Requires \usepackage{graphicx}
1416: % \includegraphics[width=400pt]{range}\\
1417: \includegraphics[width=400pt]{fig1}\\
1418: \caption{}
1419: \label{range}
1420: \end{figure}
1421: \newpage
1422:
1423: \begin{figure}[H]
1424: % % Requires \usepackage{graphicx}
1425: % \includegraphics[width=400pt]{valid_k_lacis_oinas1991}\\
1426: \includegraphics[width=400pt]{fig2}\\
1427: \caption{}
1428: \label{kdislacis}
1429: \end{figure}
1430: \newpage
1431:
1432: \begin{figure}[H]
1433: % % Requires \usepackage{graphicx}
1434: % \includegraphics[width=400pt]{valid_k_mlawer1997}\\
1435: \includegraphics[width=400pt]{fig3}\\
1436: \caption{}
1437: \label{kdismlawer}
1438: \end{figure}
1439: \newpage
1440:
1441:
1442: \begin{figure}[H]
1443: % % Requires \usepackage{graphicx}
1444: %\flushleft{
1445: % \includegraphics[width=350pt]{mrac}
1446: % \includegraphics[width=350pt]{mrac_100}\\
1447: % \includegraphics[width=470pt]{mrac_v}
1448: \includegraphics[width=470pt]{fig4}
1449: \caption{}
1450: \label{MRACvalid}%}
1451: \end{figure}
1452:
1453: \newpage
1454:
1455:
1456: \begin{figure}[H]
1457: % % Requires \usepackage{graphicx}
1458: % \includegraphics[width=400pt]{interpolation_error}\\
1459: \includegraphics[width=400pt]{fig5}\\
1460: \caption{}
1461: \label{limitstable}
1462: \end{figure}
1463:
1464: \newpage
1465:
1466: \begin{figure}[H]
1467: % % Requires \usepackage{graphicx}
1468: % \includegraphics[width=400pt]{fluxdiff}\\
1469: \includegraphics[width=400pt]{fig6}\\
1470: \caption{}
1471: \label{fluxdiff}
1472: \end{figure}
1473:
1474:
1475:
1476: \newpage
1477:
1478: \begin{figure}[H]
1479: % % Requires \usepackage{graphicx}
1480: % \includegraphics[width=400pt]{vartau}\\
1481: \includegraphics[width=400pt]{fig7}\\
1482: \caption{}
1483: \label{vartau}
1484: \end{figure}
1485:
1486:
1487: \newpage
1488:
1489: \begin{figure}[H]
1490: % Requires \usepackage{graphicx}
1491: % \includegraphics[width=400pt]{reldif}\\
1492: \includegraphics[width=400pt]{fig8}\\
1493: \caption{}
1494: \label{relhumdif}
1495: \end{figure}
1496:
1497: \newpage
1498:
1499:
1500: \begin{figure}[H]
1501: % Requires \usepackage{graphicx}
1502: % \includegraphics[width=400pt]{tprof}\\
1503: \includegraphics[width=400pt]{fig9}\\
1504: \caption{}
1505: \label{tprofile}
1506: \end{figure}
1507:
1508: \newpage
1509:
1510:
1511: \begin{figure}[H]
1512: % Requires \usepackage{graphicx}
1513: % \includegraphics[width=400pt]{wprof}\\
1514: \includegraphics[width=400pt]{fig10}\\
1515: \caption{}
1516: \label{wprofile}
1517: \end{figure}
1518:
1519: \newpage
1520:
1521:
1522: \begin{figure}[H]
1523: % Requires \usepackage{graphicx}
1524: % \includegraphics[width=400pt]{compkasting_paper}\\
1525: \includegraphics[width=400pt]{fig11}\\
1526: \caption{}
1527: \label{minin2data}
1528: \end{figure}
1529:
1530:
1531: \end{document}
1532:
1533:
1534: Despite a fainter Sun, the surface of the early Earth was mostly
1535: ice-free. Proposed solutions to this so-called "faint young Sun
1536: problem" have usually involved higher amounts of greenhouse gases
1537: than present in the modern-day atmosphere. However, geological
1538: evidence seemed to indicate that the atmospheric CO2 concentrations
1539: during the Archaean and Proterozoic were far too low to keep the
1540: surface from freezing. With a radiative-convective model including
1541: new, updated thermal absorption coefficients, we found that the
1542: amount of CO2 necessary to obtain 273 K at the surface is reduced up
1543: to an order of magnitude compared to previous studies. For the late
1544: Archaean and early Proterozoic period of the Earth, we calculate
1545: that CO2 partial pressures of only about 2.9 mb are required to keep
1546: its surface from freezing which is compatible with the amount
1547: inferred from sediment studies. This conclusion was not
1548: significantly changed when we varied model parameters such as
1549: relative humidity or surface albedo, obtaining CO2 partial pressures
1550: for the late Archaean between 1.5 and 5.5 mb. Thus, the
1551: contradiction between sediment data and model results disappears for
1552: the late Archaean and early Proterozoic.
1553:
1554: Planetary and Space Science, 2008, vol. 56, p. 1244-1259, ISSN:
1555: 0032-0633
1556:
1557:
1558: 53 pages, 4 tables, 11 figures, published in Planetary and Space
1559: Science
1560:
1561: 0804.4134.philip_von_paris.23136 was accepted. Article-id:
1562: 0804.4134, Article password: seecm (access no longer password
1563: restricted)
1564: