• Keine Ergebnisse gefunden

Regionally Resolved Diagnostic of Transport: A Simplified Forward Model for CO2

N/A
N/A
Protected

Academic year: 2022

Aktie "Regionally Resolved Diagnostic of Transport: A Simplified Forward Model for CO2"

Copied!
12
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Regionally Resolved Diagnostic of Transport: A Simplified Forward Model for CO

2

PAULKONOPKA, FELIXPLOEGER, MENGCHUTAO,ANDMARTINRIESE IEK-7, Forschungszentrum Julich, J ulich, Germany

(Manuscript received 20 December 2016, in final form 2 June 2017) ABSTRACT

Simply diagnostic tools are useful for understanding transport processes in complex chemistry transport models (CTMs). For this purpose, a combined use of the airmass origin fractions (AOFs) and regionally resolved mean ages (RMAs) is presented. This approach merges the concept of the origin of air with the well- known theory of the mean age of air (AoA) for different regions covering the whole Earth. The authors show how the AoA calculated relative to Earth’s surface can be decomposed into regionally resolved components (i.e., into RMAs). Using both AOFs and RMAs, the authors discuss differences in the seasonality of transport from the Northern and Southern Hemispheres into the tropical tropopause layer (TTL), the asymmetries of the interhemispheric exchange, and differences in relation to the continental or oceanic origin of air.

Furthermore, a simplified transport model for a chemically passive species (tracer) is formulated that has some potential to approximate the full transport within a CTM. This analytic approach uses the AOFs as well as the RMAs as parameters to propagate a tracer prescribed on Earth’s surface (lower boundary condition).

This method is exactly valid for sources that change linearly with time in each of the considered regions. The authors analyze how well this approach approximates the propagation of CO2from the planetary boundary layer (PBL) into the whole atmosphere. The CO2values in the PBL are specified by the CarbonTracker dataset. The authors discuss how this approach can be used for inverse modeling of CO2.

1. Introduction

Although tropospheric air enters the stratosphere pre- dominantly through the tropical tropopause layer (TTL) (Fueglistaler et al. 2009), there is a wide range of pathways connecting the planetary boundary layer (PBL) with the TTL itself (e.g.,Levine et al. 2007;Randel et al. 2010;

Vogel et al. 2011). Different diagnostics have been de- veloped to quantify such pathways. The age spectrum of air has been shown to be a comprehensive diagnostic of atmospheric transport defining the composition of air as a probability density function (PDF) of transit times from a prescribed source region to the considered space–time point (e.g.,Hall and Plumb 1994;Waugh and Hall 2002;

Orbe et al. 2015,2016).

In addition to the assumption that chemistry does not affect the transported (idealized) species (i.e., the tracers), the definition of the source region is another limiting factor in determining the age spectrum. Typically, the tropical

tropopause is considered as a source region if only strato- spheric transport is envisaged. Another useful choice is to consider Earth’s surface, relative to which the PDF of the transit times has to be determined. In such a case, transport within the troposphere can also be diagnosed.

However, it might also be useful to compare transport from different parts of Earth, for example, to contrast the contributions of the oceans relative to the continents or of the northern relative to the Southern Hemisphere (Holzer 2009;Orbe et al. 2015;Tissier and Legras 2016).

Furthermore, if emissions such as carbon dioxide (CO2) are considered, it can also be useful to quantify transport from different continents or even countries. Although a regionally resolved diagnostic of transport in terms of age spectra is a natural generalization (Holzer and Boer 2001), its implementation would require large computer resources as the calculation of age spectra for only one source region is a demanding and resource-intensive task (Haine et al. 2008;Ploeger and Birner 2016).

Commonly, only parts of the age spectrum can be de- rived from measurements, typically those describing the

Corresponding author: Paul Konopka, p.konopka@fz-juelich.de Denotes content that is immediately available upon publica- tion as open access.

This article is licensed under aCreative Commons Attribution 4.0 license (http://creativecommons.

org/licenses/by/4.0/).

DOI: 10.1175/JAS-D-16-0367.1

Ó2017 American Meteorological Society

(2)

fast transport to the considered space–time point and, consequently, quantifying the contribution of young air within the age spectrum (Ehhalt et al. 2007). Furthermore, the first moment of the age spectrum [i.e., the mean age of air (AoA) (Hall and Plumb 1994)] can also be derived from the observations and is widely used, for example, to quantify differences in stratospheric transport among comprehensive climate models (Eyring et al. 2010).

Recently,Orbe et al. (2013,2015) introduced a cli- matology of mass fractions of air originating from dif- ferent regions on Earth and entering the stratosphere.

As we discuss later, such mass fractions can also be un- derstood as zero moments of the regional age spectra. In addition,Waugh et al. (2013)andOrbe et al. (2016)also discussed the seasonality of the mean age from the Northern Hemisphere (NH), which is an example of the regionally resolved mean age. Following the spirit of these investigations, we divide the whole Earth into different regions and in this paper analyze how transport can be diagnosed in terms of the airmass origin fractions (AOFs) and the regionally resolved mean ages (RMAs).

First, we investigate the differences in the seasonality of the air composition in the TTL with respect to the origin of air. Here, interhemispheric differences as well as differ- ences between continents and oceans will be considered. In particular, we discuss how the common AoA calculated relative to the whole PBL can be decomposed into its re- gional components. Second, we show that by using both AOFs and RMAs, an arbitrary tracer with specified line- arly growing sources in every region can be analytically reconstructed at every space–time point in the atmosphere.

Although such a reconstruction strictly applies only for linear sources (see next section), we use this method to determine the global CO2distribution with specified realistic spatial distributions and time evolution of the sources in the PBL as derived from the CarbonTracker dataset (Peters et al. 2007). We show that such a sim- plified forward model reproduces fairly well the CO2 distribution derived from a full chemistry transport model (CTM), especially if only the total column of CO2

is considered.

In the next section, we derive all necessary equations.

Sections 3and4discuss the regionally resolved seasonality of transport from the PBL to the TTL by extending the results presented in Orbe et al. (2015). Insection 5, the simplified forward model of CO2 transport is presented.

Finally, we discuss our results insection 6.

2. Airmass origin fractions and regionally resolved mean ages

The formal solution of the general transport equation for the mixing ratio of a conserved and passive tracer

m that is prescribed at the boundary V (e.g., Earth’s surface) can be written in terms of an age spectrumG, also called transit-time distribution (Hall and Plumb 1994;Waugh and Hall 2002):

m(r,t)5 ð

0

m(V,t2t)G(r,V,t,t2t)dt, (1) where m(V,t2t) is the tracer mixing ratio at the boundary surfaceVtimetago.

Now, we decompose the boundary V into a set of subregions Vi with

å

Vi5 V, i51,. . .,n, and with specified time seriesmi(t) in each regionVi and 0 else- where (seeFig. 1) such that

m(V,t)5

å

n

i51mi(t) . (2) In addition, definingGiinsideViand 0 elsewhere allows us to write (Haine and Hall 2002)

G(r,V,t,t0)5

å

n

i51

Gi(r,V,t,t0), (3)

and, consequently, we can rewrite relation(1)as m(r,t)5

å

n

i51

ð

0

mi(t2t)Gi(r,t,t2t)dt,

Where, from now on, we do not write theVdependence explicitly.

The term G(r,t,t0) can also be understood as a probability distribution function as it is the mass fraction of an air parcel that has left the boundary surface V time t ago, hence a transit-time PDF

FIG. 1. The decomposition of the boundaryV (such as that of Earth’s surface) into thensubregions (such as continents and oceans), in which the passive tracer is specified in the form of time seriesmi(t).

(3)

(e.g., Hall and Plumb 1994; Schoeberl et al. 2000).

This means thatGis normalized to 1; that is, ð

0

G(r,t,t2t)dt51 . (4) With the special choice mi(t)5H(t2t0) (with H being the Heaviside step function) within each region Vi, the AOFsfi,i51,. . .,ncan be defined as (Orbe et al. 2013)

fi(r,t):5 lim

t0/2‘

ð

0

H(t2t02t)Gi(r,t,t2t)dt 5

ð

0

Gi(r,t,t2t)dt. (5) Because of the normalization condition(4), the sum of the fractions equals 1 at every space–time point;

that is,

å

n

i51fi(r,t)51 . (6)

Note that such idealized origin tracers can be simply implemented into a CTM by settingfito 1 inside Vi

and 0 elsewhere inV.

Furthermore, properly normalized transit-time PDFs for the individual source regions Vi can be defined as Gi/fi. Their first moment over transit timet

Gi(r,t)51 fi

ð

0

tGi(r,t,t2t)dt (7)

defines the mean age of air with respect to the source regionVior simply RMAs.

In particular, for a linear source functionmi, with mi(t)5a1bt, a,b5const , (8) the relation

mi(r,t)5fi(r,t)mi[t2 Gi(r,t)] (9) holds (for the derivation see below); that is, the time delay between the concentration mi at the space–time point (r,t) and the linearly increasing concentration at the source mi is equal to the mean ageGi.

The linearity ofmiand the normalization condition ofGirender the derivation of(9)very simple. Using the same arguments as in Hall and Plumb (1994), we obtain for the space–time distribution of the tracermi:

mi(r,t)5 ð

0

mi(t2t)Gi(r,t,t2t)dt 5

ð

0

[a1b(t2t)]Gi(r,t,t2t)dt

5fi(r,t)[a1bt2bGi(r,t)]5fi(r,t)mi[t2 Gi(r,t)].

It is also relatively easy to derive the spectrum of Gi from a CTM. For this purpose, a linearly increasing tracer, as described above, has to be defined in eachViregion and set to zero elsewhere in the boundary layer. Thus, for the sum of the linear source functionsmi[see Eq.(8)], we can write

m(r,t)5

å

n

i51mi(r,t)5

å

n

i51

fi(r,t)mi[t2 Gi(r,t)], (10) where relation(9)was used. Because of

m(r,t)5a1b[t2 G(r,t)]

5

å

n

i51fi(r,t)fa1b[t2 Gi(r,t)]g (11)

the following relation connectsGwith theGi: G 5

å

n

i51fiGi. (12)

Finally, after initializing a CTM covering the troposphere and stratosphere, a spinup period of aroundm510,. . ., 20 years is expected until equi- librium with the boundary conditions has been reached (Orbe et al. 2013,2015), such that all thefi

andGido not depend on the initialization timet0: fi(r,t,t0),Gi(r,t,t0)t.15 yr! fi(r,t),Gi(r,t). (13) A convenient way to achieve such a quasi- stationary state is to repeat the first year of the simulationmtimes. Within such a perpetuum run, all species at the end of the year are interpolated on the first day of the same year. Although there is a discontinuous jump in the flow field due to such interpolation, this approach can significantly reduce nu- merical costs if Lagrangian CTMs are used [like Chemical Lagrangian Model of the Stratosphere (CLaMS) in this study; see next section]. In such a case, the same trajec- tories as calculated for the first year of the perpetuum run can also be used in the followingmyears, reducing the total numerical costs by more than 80%.

In particular, the linearly increasing tracersmihave to be redefined at the end of each perpetuum period using thefi[see relation(5)]. Because of

(4)

m(r,t2 Dt)5m(r,t)2bDt5

å

n

i51[mi(r,t)2fi(r,t)bDt]

5

å

n

i51mi(r,t2 Dt) (14) withDtdenoting the perpetuum time period (here: 1 yr), we obtain the relation

mi(r,t2 Dt)5mi(r,t)2fi(r,t)bDt. (15)

3. Diagnostic of transport into the TTL: An example

As a CTM we use the 2005 perpetuum run of CLaMS (McKenna et al. 2002;Konopka et al. 2004;Pommrich et al. 2014) driven by the horizontal winds and diabatic heating rates (vertical velocities) derived from ERA- Interim (Dee et al. 2011). To resolve transport pro- cesses in the troposphere influenced by the orography and transport processes in the stratosphere where adiabatic horizontal transport dominates, the hybrid s–ucoordinatezis used (wheresis a terrain-following coordinate anduis potential temperature) (Mahowald et al. 2002;Konopka et al. 2007). More details can be found inPommrich et al. (2014).

All idealized tracers [i.e., those from which the fi

(AOFs) andGi(RMAs) can be calculated] are set in the orography-following lower boundary of CLaMS de- fined by the condition 0#z#100 K (i.e., within a layer covering roughly the first 1.5 km of the atmosphere above the ground). In the following, we consider this layer of the CLaMS air parcels as the simplest approximation of the PBL. At the upper boundary, all the air parcels in the layer between z 5 2200 and 2500 K (around 0.1 hPa) are replaced by their initial configuration (same area of each air parcel depending on the resolution used) and all tracers are interpolated from their horizontal next neighbors in this layer (open boundary condition).

The 34 yr of CLaMS simulation of CO2(from 1979 to 2012) extend between Earth’s surface and the mesosphere. Until the year 2000, a zonally symmetric lower boundary condition is used, which is derived from the NOAA/Climate Monitoring and Di- agnostics Laboratory (CMDL) ground-based mea- surement network (Masarie and Tans 1995;Novelli et al. 2003) as described inPommrich et al. (2014). For the 2000–12 period, the zonally resolved lower boundary was derived from the CarbonTracker dataset (Peters et al. 2007). Here the first five lowest levels of the CarbonTracker dataset (run CT2013B, available every 3 h; seehttps://www.esrl.noaa.gov/gmd/ccgg/

carbontracker/), which roughly cover the PBL resolved by CLaMS, were vertically averaged and included in the CLaMS simulation as a 2D (longitude–latitude) lower boundary condition. In the following, the CO2time–space evolution derived from such a CLaMS simulation is used as a refer- ence—that is, as a proxy of a ‘‘true’’ CO2distribution in the atmosphere.

To diagnose transport, in particular to determine the fiandGi, CLaMS perpetuum runs for 2005 are carried out with a horizontal model resolution of 100 km (mean horizontal distance between air parcels) and vertical resolution of;400 m around the tropopause.

For sensitivity studies, results with reduced horizontal and vertical resolutions of 200 km and 800 m are shown. After each year of the 40-yr perpetuum run, the AOFs and RMAs are interpolated from the last day (31 December) to the first day (1 January) of the considered year whereas relation (15)is used for the linearly increasing tracers.

As shown later, we ensure in this way that the equilibrated state was reached; that is, allfiandGido not depend on the initialization time t0. Note that our way of calculating AOFs and RMAs from a per- petuum run (instead of from a sufficiently long tran- sient run) neglects some interannual variability of transport such as that due to the El Niño–Southern Oscillation (ENSO) or due to the quasi-biennial os- cillation (QBO). Especially air masses older than 1 yr (perpetuum cycle) are affected by our approximation, although they are of minor importance for our analysis (shown later).

An example of AOFs and RMAs, calculated for 15 December 2005, can be seen in Fig. 2. Here, 36 almost-equal source regions are defined as a regular grid covering Earth’s surface (black grid; note that boxes covering the polar caps are slightly larger than all other boxes in the tropics and extratropics). The TTL is considered as a destination region over which allfiandGiare averaged (thick black box). For better comparison withOrbe et al. (2015), the TTL is defined as a domain between 208S and 208N and between 100–70 hPa.

As expected, the largest fraction of air in the TTL originates from the tropical belt between 208S and 208N (in total more than 60%) with the highest contribution from the Maritime Continent, tropical Pacific, and Indian Ocean. On the other hand, the youngest air (i.e., the fastest mean transport) occurs over the tropical/subtropical Pacific, whereas the old- est air originates from the southern polar region. Note that although the highest values of AOFs are located around the equator there is a stronger hemi- spheric asymmetry in the distribution of RMAs with

(5)

much older air from the southern than from the northern high latitudes.

It is important to note that in order to obtain results that are independent of the initialization time of the artificial tracers, a quasi-stationary state has to be reached. As an example,Fig. 3shows how the values off10 andG10 (tenth region within 08–208N, 08–608E;

red dashed box inFig. 2) converge to such independent values after 40 yr of the perpetuum simulation (note that f10andG10converge to their final values after;8 and;15 perpetuum years, respectively). Then, initial conditions can be ruled out andfiandGican be understood as di- agnostics of pure transport.

4. Seasonality of regionally resolved mean ages in the TTL

First, we compare our results for AOFs with similar results published byOrbe et al. (2015)(although a dif- ferent CTM and different winds were used) and define

the source and destination regions as shown in Fig. 4.

FollowingOrbe et al. (2015), three source regions in the PBL (0–1.5 km) are considered: Southern Hemisphere (SH; 908–108S; green), tropics (108S–108N; red), and the Northern Hemisphere (NH; 108–908N; cyan). Beyond the TTL as a destination region, we also consider two additional regions in the extratropical troposphere (TR): TR–SH (800–700 hPa, 908–508S) and TR–NH (800–700 hPa, 508–908N).

The seasonality of the respectivefiandGifor all three destination regions (TTL, TR–SH, and TR–NH) is shown in Fig. 5. The top panel of Fig. 5a shows the seasonalities of the mass fractionsfi, which are roughly the same as those discussed inOrbe et al. (2015) (see their Fig. 2a); that is, the composition of the TTL is dominated (by more than ;60%) by the air from the tropical belt between 108S and 108N and with a maxi- mum (minimum) around June or July (November).

The tropical belt also determines the mean age in the TTL as can be inferred from the bottom panel ofFig. 5a:

according to Eq. (12), the mean age of the total Earth’s surface (black) can be decomposed into the tropical con- tribution (red) as well as the extratropical contributions of the Southern (green) and Northern (blue) Hemispheres.

Note that the black lines and the red lines are very similar, indicating that the mean age relative to the total Earth’s surface is almost the same as the mean age relative to the tropical belt. It should be mentioned that the youngest air from the extratropical Southern (Northern) Hemisphere is expected around April (December) with a respective mass fraction of;20% (;40%). The higher contribution of the extratropical air from the NH during December than from the SH during April is a result of the delayed influence of the Hadley cell with the equatorward branch being stronger during the summer in the NH than during the winter in the SH.

Now, we perform the same type of analysis for the polar tropospheric air between 700 and 800 hPa (see TR–SH and TR–NH regions inFig. 4). The respective seasonalities offiandGi(Figs. 5b and 5c) quantify fairly well the asymmetries in hemispheric and the inter- hemispheric transport. First, tropical air (red) reaches the TR–NH region more effectively than the TR–SH region with the fastest transport in the winter hemi- sphere because of more effective quasi-isentropic stir- ring by tropospheric eddies during the winter than during the summer and in the NH than in the SH.

Furthermore, the mass fraction of the NH PBL in the SH polar troposphere amounts to ;6% [blue curve in Fig. 5b(top)]. The reverse contribution (i.e., that of the SH PBL in the NH polar troposphere) is roughly 3 times smaller [green curve inFig. 5c(top)]. The comparison of the regional mean ages reveals that, on average, the

FIG. 2. Regionally resolved mass fractions (AOFs) and mean ages of air (RMAs) within the TTL defined as a domain between 208S and 208N and between 100 and 70 hPa (thick black region) are calculated as an example for 15 Dec 2005. In total, 36 equal source regions are considered (thin black grid). Thus, 36 color-coded values of AOFs and RMAs regionally resolve the composition (mass fractions) as well as the mean ages of transport from the PBL to the TTL. For the selected box (dashed red), the convergence of the AOF and RMA values during the perpetuum integration is shown inFig. 3.

(6)

fastest transport from the NH PBL to the SH polar troposphere can be diagnosed around October and vice versa around May.

The values of the mean age for the transport from the extratropical NH to the polar SH troposphere can also

be compared with the mean age derived from the SF6 observations (Waugh et al. 2013). As SF6 is emitted predominantly in NH midlatitudes and its growth rate is nearly linear, the time lag between the SF6mixing ratio at a given location and that at the NH midlatitude

FIG. 4. Definition of three zonally symmetric and orography-following source regions for the PBL (0–1.5 km) for the Southern Hemisphere (908–108S; green), tropics (108S–108N; red), and the Northern Hemisphere (108–908N; cyan) (Orbe et al. 2015). As the destination regions (i.e., regions where transport will be diagnosed) the following three zonally symmetric domains are considered: TTL (100–70 hPa, 208S–208N), TR–SH (800–700 hPa, 908–508S), and TR–NH (800–700 hPa, 508–908N).

FIG. 3. Results of a perpetuum simulation forf10andG10(tenth region within 08–208N, 08–608E; dashed red box inFig. 2). Red values are reached after 40 perpetuum years for 2005. After about 8 and 15 yr, respectively, the AOFs and the RMAs have equilibrated; that is, the normalization conditions in(6)and(12)are valid and all the results are independent of the initialization time. The TTL (destination region inFig. 2) is also schematically labeled.

(7)

surface provides an estimate of the mean age from the NH surface.Waugh et al. (2013)report values of around 16–18 months for the AoA of SF6 at 608S (see their Fig. 3), which roughly correspond to our estimate of around 15 months (see blue line inFig. 5b) (bottom).

However, the seasonality of AoA in TR–SH is very similar to that derived from SF6observations at Cape Grim (Tasmania, 40.78S) and at the South Pole (Antarctica, 908S) as described inWaugh et al. (2013).

Note that the analysis of AOFs and RMAs does not allow any conclusions to be drawn about the possible pathways of transport.Orbe et al. (2016)have recently shown that the transit-time distribution (TTD) (i.e., the full age spectrum) provides much more detailed information on transport from the NH midlatitude surface. In particular, huge differences in the season- ality of the pathways were diagnosed although the respective differences in the mean ages were much smaller. It should be also emphasized that all the di- agnostic results discussed here and in other studies strongly depend on the quality of the wind fields used to run the CTM and how unresolved processes such as convection, gravity waves, or mixing are parameterized in the model.

Finally, the same type of analysis, but for the source regions defined by all continents and all oceans on

Earth, is shown in Fig. 6. In the top panel, only the mass fraction of the continents f (red) is shown be- cause the contribution of the oceans (12f) follows from the normalization condition. The dashed line denotes the geometric percentage of continents of 29.7%. Thus, only during SON is the contribution of the continents is slightly higher (2%–3%) than a simple geometric estimate. The analysis of respective

FIG. 6. (top) Mass fractions and (bottom) the respective mean ages for the continent and ocean source regions. The geometric percentage of continents (29.7%) is also shown.

FIG. 5. Seasonality of (top) the mass fractionsfiand (bottom) the mean agesGifor different source regions (colors) and for three destination regions defined inFig. 4: (a) TTL, (b) TR–SH, and (c) TR–NH.

(8)

mean ages shows that the mean age of the continental/

oceanic air only differs slightly from the mean age calculated relative to the whole Earth’s surface.

Furthermore, the strongest contribution of the con- tinents during the SON season is accompanied by the fastest mean transport if compared with the oceans, although the differences are less than 1 month. All these seasonalities depend only weakly on the cho- sen year (here 2005) or on the method how the equilibrated state was reached (perpetuum versus transient run), as confirmed for the CLaMS low- resolution case.

5. A simplified forward model

Now we apply relation(10) to reconstruct the tem- poral evolution of the CO2 distribution in the whole atmosphere. The use of relation (10) can also be un- derstood as a (simplified) forward model propagating CO2from the PBL into the whole atmosphere. Because we can derive the true CO2 distribution from a CTM (here CLaMS) and compare this distribution with that derived from Eq.(10), the quality of this approximation can be verified in this way.

The regions are defined either by a regular longitude–

latitude (Xn/Ym) grid with roughly the same area of every (n,m)th grid cell or by defining different areas for continents and oceans (Cn/Om, where n and m are the numbers of continental and oceanic regions,

respectively), known as nested grids (seeFig. 7). Thus, the following grids will be used: X6/Y6, X12/Y12, C1/O1, C17/O18, and C78/O18 with 36, 124, 2, 35, and 96 regions, respectively.

Thus, using relation(10),m(r,t) denotes a CO2mixing ratio that is reconstructed from the mixing ratiosmiin the PBL, which are specified over the elapsed time and averaged over each regioni. Furthermore, thefiandGi

contain information about the amount and time scale of tracers transported from the regionito the considered space–time (r,t) and are derived from the 2005 perpetuum run with CLaMS.

An example of such a reconstruction (calculated for 15 December 2006 by using the C78/O18 grid) is shown in the right-hand column ofFig. 8. The left-hand column shows the true CO2 distribution as derived from the full CLaMS simulation without dividing the PBL into regions.

In particular, in the top two panels, the CO2mixing ratio in the PBL (i.e., in the lowest layer of CLaMS) can be seen with values derived from the CarbonTracker dataset (left) and after averaging this field over the regions according to the relation (10). Note that mi[t2 Gi(r,t)] in Eq.(10)denotes CO2mixing ratios at the surface shifted to the past by the mean ageGiat the considered space–time point (r,t) and averaged over the region i. Because we cannot show all the PBL values in the past, here only the values for 1200 UTC 15 December 2006 are plotted.

FIG. 7. (left) Two types of grids are used: a regular one with roughly equal surface for each box (black) and nested grids that can resolve continents with a higher resolution than the oceans (red). (right) An example with 17 and 18 regions covering continents and oceans, re- spectively (C17/O18). A regular distribution of the Lagrangian air parcels is shown in the lowest model layer with;1.5-km thickness for a low-resolution CLaMS configuration (horizontal resolution around 200 km).

(9)

FIG. 8. (left) Full transport of CO2with CLaMS vs (right) simplified transport following relation(10). (top) Specified CO2distribution in the PBL (left) as derived from the CarbonTracker dataset and (right) as averaged over the grid boxes in order to obtain input for the simplified transport according to Eq.(10). (top middle),(bottom middle) Transported CO2as derived (left) directly from CLaMS and (right) by the use of the forward model(10).

The parametersfiandGiare calculated from a 15-yr perpetuum run with CLaMS. (bottom) The column-averaged CO2(XCO2) is a proxy for the data that can be expected from a satellite. All panels are examples for 1200 UTC 15 Dec 2006.

(10)

The second and third rows in Fig. 8 depict the re- spective snapshots (i.e., for the same date) in two dif- ferent pressure layers (600–700 and 120–150 hPa). Once again, whereas in the left-hand column, the CO2mixing ratios in CLaMS air parcels are calculated from the full Lagrangian transport (advection 1mixing), the right- hand column uses Eq.(10)with precalculated AOFsfi and RMAsGi. In the bottom row, the column-averaged XCO2is shown first by intepolating the irregular grid of Lagrangian air parcels on a regular pressure grid and then using the relation

XCO25 ð

0

NCO

2

(z)dz ð

0

Nair(z)dz

(16)

withNCO2(z) andNair(z) denoting the number density of CO2and of air, respectively (zis log-pressure altitude).

Note that although differences in the CO2distribution in the PBL used in both procedures are significant (cf.

top panels ofFig. 8), the differences in the respective XCO2are much smaller.Figure 9depicts such relative differences in XCO2derived from CLaMS simulations with the highest nested resolution of the regions (C78/

O18). We conclude that by using Eq.(10)the differences in XCO2relative to a resolved full transport are smaller than 1% (i.e., they are on the order of the accuracy of current and future satellite instruments measuring XCO2). It is also worth noting that a sharp edge around 208N in the relative differences shown inFig. 9is a result of the definition of the regions (see top-right panel in Fig. 8), which has still some potential for improvement.

Finally, we discuss how the performance of our ap- proach depends on the grid resolution and on the choice of the mean ages Gi. In the left panel of Fig. 10, the vertically resolved differences (root-mean-square de- viations in every CLaMS layer) between the results of the full and simplified transport calculations are shown.

In the left panel, only one mean ageGcalculated for the

FIG. 9. Relative difference in the reconstructed and full column of CO2as calculated for the C78/O18 model configurations on 15 Dec 2005.

FIG. 10. Vertically resolved cumulative error (root-mean-square deviation in every model layer) for different grid configurations (colors; thickness) (left) by using only one mean age of air for the whole surface of EarthGand (right) by applying all regionalGi.

(11)

whole surface of Earth is used in Eq.(10)(i.e.,Gi5 G,

i51,. . .,n), whereas in the right panel, all the regional

Giare taken into account.

It can be seen that below 500 hPa the C78/O18 model configuration (105 regions) performs the best with mean differences smaller than 1.5 ppmv. The use of regionally resolvedGiimproves the CO2reconstruction by more than 50%. Note that taking into account more continental regions than oceanic regions helps to keep numerical efforts low (105 regions of the C78/O18 configuration versus 124 regions of the X12/Y12 regu- lar grid) and also reduces the errors below 500 hPa by more than 50%. By using a transient instead of a per- petuum run (case C17/O18), the results shown in Fig. 10 are exactly the same below 200 hPa with no significant improvement above this level, consistent with transit times usually smaller than 1 yr below this level (not shown).

6. Conclusions

Starting from the airmass origin fractions (AOFs) defining the contribution of different regions on Earth to the composition of air at a given space–time point, we extended this analysis by considering the re- gionally resolved mean ages (RMAs) characterizing the mean transport from the regions to the space–time point considered. We showed how this additional di- agnostic allows a more detailed characterization of transport.

In particular, the mean age of air (AoA), calculated relative to the PBL of the whole Earth, can be decom- posed into its regionally resolved components(12). For example, the seasonality of the AoA in the TTL is the same as for the air originating in the tropical belt be- tween 108S and 108N (oldest air around August or September,;60% contribution) whereas the season- alities of the air originating south of 108S (north of 108N) peak 1–2 months earlier (later), both with a contribution of ;20%. In addition, the asymmetry of interhemispheric exchange could be quantified showing a faster mean transport from the SH and tropics into the lower NH polar troposphere than vice versa. Furthermore, the seasonality of mean ages from the continents and oceans is very similar with the youngest air from December to May and the oldest air from June to September.

We have also shown how knowledge of the time evolution of the AOFs and RMAs at every grid point of the model allows us to reconstruct a passive tracer with known linear sources everywhere in the atmosphere.

The AOFs and RMAs can be easily calculated from the pure transport part of a CTM (i.e., the chemistry

modules can be switched off). Typically CTMs are driven by reanalysis winds (here ERA-Interim), which extend over a few decades and make the perpetuum runs (as used in this paper) only necessary for the first year of the simulation. Furthermore, the CTM has to be run only once to determine the transport properties (i.e., the AOFs and RMAs) and then the simplified forward model can be applied to any tracer given its concentra- tion at the surface.

Such a simplified model was applied and validated by studying CO2propagation from Earth’s surface into the atmosphere extending up to the stratopause. Currently, using satellite observations such as those of the Euro- pean Scanning Imaging Absorption Spectrometer for Atmospheric Chartography (SCIAMACHY), Japan’s Greenhouse Gases Observing Satellite(GOSAT), or the recently launched U.S. Orbiting Carbon Observatory (OCO), only the column of CO2or XCO2(i.e., the mean mixing ratio of CO2averaged over the column) can be derived. The differences between the full and simplified transport, especially in terms of column-averaged XCO2, are smaller than 1% (i.e., on the order of the accuracy of the satelliteborne measurements).

Thus, the forward model presented offers a simple framework for the inverse modeling of CO2also known as the top-down approach (Nisbet and Weiss 2010). In particular, using this approach, the sensitivity of XCO2

with respect to the distribution of CO2sources, such as their spatial inhomogeneities or their diurnal and sea- sonal cycles, can be analyzed in order to find out which properties of XCO2 variability reflect anthropogenic footprints. A conceptually similar type of analysis was discussed inStohl (2006)where the sensitivity of Arctic pollution to different midlatitude source regions was quantified in terms of the Arctic age (i.e., the time air resides continuously north of 708N).

Acknowledgments.The European Centre for Medium- Range Weather Forecasts (ECMWF) provided meteoro- logical analysis for this study. We thank Ann-Sophie Tissier and Bernard Legras for providing us with nu- merical support for orographic definitions of the conti- nental regions. Excellent programming support was provided by N. Thomas. The authors sincerely thank Mohamadou Diallo for support related to the Carbon- Tracker data. We also thank Rolf Müller for helpful discussions. Finally, we thank all three reviewers for their insightful and probably very time-consuming re- views, as these comments led us to an improvement of the work. This research was supported by the German Helmholtz-Gemeinschaft within the Helmholtz-CAS joint research group (JRG) ‘‘Climatological impact of increasing anthropogenic emissions over Asia.’’

(12)

REFERENCES

Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis:

Configuration and performance of the data assimilation system.Quart. J. Roy. Meteor. Soc.,137, 553–597, doi:10.1002/

qj.828.

Ehhalt, D. H., F. Rohrer, D. R. Blake, D. E. Kinnison, and P. Konopka, 2007: On the use of nonmethane hydrocar- bons for the determination of age spectra in the lower stratosphere. J. Geophys. Res., 112, D12208, doi:10.1029/

2006JD007686.

Eyring V., T. G. Sheperd, and D. W. Waugh, Eds., 2010: SPARC CCMVal report on the evaluation of chemistry-climate models. SPARC Rep. 5, WCRP 30, WMO/TD 40, 434 pp.

[Available online athttp://www.sparc-climate.org/publications/

sparc-reports/sparc-report-no5/.]

Fueglistaler, S., A. E. Dessler, T. J. Dunkerton, I. Folkins, Q. Fu, and P. W. Mote, 2009: Tropical tropopause layer. Rev.

Geophys.,47, RG1004, doi:10.1029/2008RG000267.

Haine, T. W. N., and T. M. Hall, 2002: A generalized transport theory: Water-mass composition and age.J. Phys.

Oceanogr.,32, 1932–1946, doi:10.1175/1520-0485(2002)032,1932:

AGTTWM.2.0.CO;2.

——, H. Zhang, D. W. Waugh, and M. Holzer, 2008: On transit time distributions in unsteady circulation models. Ocean Modell.,21, 35–45, doi:10.1016/j.ocemod.2007.11.004.

Hall, T. M., and R. A. Plumb, 1994: Age as a diagnostic of strato- spheric transport.J. Geophys. Res.,99, 1059–1070, doi:10.1029/

93JD03192.

Holzer, M., 2009: The path density of interhemispheric surface- to-surface transport. Part I: Development of the diagnostic and illustration with an analytic model. J. Atmos. Sci.,66, 2159–2171, doi:10.1175/2009JAS2894.1.

——, and G. J. Boer, 2001: Simulated changes in atmospheric transport climate. J. Climate, 14, 4398–4420, doi:10.1175/

1520-0442(2001)014,4398:SCIATC.2.0.CO;2.

Konopka, P., and Coauthors, 2004: Mixing and ozone loss in the 1999–

2000 Arctic vortex: Simulations with the three-dimensional Chemical Lagrangian Model of the Stratosphere (CLaMS).

J. Geophys. Res.,109, D02315, doi:10.1029/2003JD003792.

——, and Coauthors, 2007: Contribution of mixing to upward transport across the tropical tropopause layer (TTL).Atmos.

Chem. Phys.,7, 3285–3308, doi:10.5194/acp-7-3285-2007.

Levine, J. G., P. Braesicke, N. R. P. Harris, N. S. Savage, and J. A.

Pyle, 2007: Pathways and timescales for troposphere-to- stratosphere transport via the tropical tropopause layer and their relevance for very short lived substances.J. Geophys.

Res.,112, D04308, doi:10.1029/2005JD006940.

Mahowald, N. M., R. A. Plumb, P. J. Rasch, J. del Corral, F. Sassi, and W. Heres, 2002: Stratospheric transport in a three- dimensional isentropic coordinate model. J. Geophys. Res., 107, 4254, doi:10.1029/2001JD001313.

Masarie, K., and P. Tans, 1995: Extension and integration of atmospheric carbon dioxide data into a globally consistent measurement record. J. Geophys. Res.,100, 11 593–11 610, doi:10.1029/95JD00859.

McKenna, D. S., P. Konopka, J.-U. Grooß, G. Günther, R. Müller, R. Spang, D. Offermann, and Y. Orsolini, 2002: A new Chemical Lagrangian Model of the Stratosphere (CLaMS) 1.

Formulation of advection and mixing.J. Geophys. Res.,107, 4309, doi:10.1029/2000JD000114.

Nisbet, E., and R. Weiss, 2010: Top-down versus bottom-up.

Science,328, 1241–1243, doi:10.1126/science.1189936.

Novelli, P. C., K. A. Masarie, P. M. Lang, B. D. Hall, R. C. Myers, and J. W. Elkins, 2003: Reanalysis of tropospheric CO trends:

Effects of the 1997–1998 wildfires.J. Geophys. Res.,108, 4464, doi:10.1029/2002JD003031.

Orbe, C., M. Holzer, L. M. Polvani, and D. Waugh, 2013: Air-mass origin as a diagnostic of tropospheric transport.J. Geophys.

Res. Atmos.,118, 1459–1470, doi:10.1002/jgrd.50133.

——, D. W. Waugh, and P. A. Newman, 2015: Air-mass origin in the tropical lower stratosphere: The influence of Asian boundary layer air. Geophys. Res. Lett., 42, 4240–4248, doi:10.1002/2015GL063937.

——, ——, ——, and S. Steenrod, 2016: The transit-time distribu- tion from the Northern Hemisphere midlatitude surface.

J. Atmos. Sci.,73, 3785–3801, doi:10.1175/JAS-D-15-0289.1.

Peters, W., and Coauthors, 2007: An atmospheric perspective on North American carbon dioxide exchange: CarbonTracker.

Proc. Natl. Acad. Sci. USA,104, 18 925–18 930, doi:10.1073/

pnas.0708986104.

Ploeger, F., and T. Birner, 2016: Seasonal and inter-annual vari- ability of lower stratospheric age of air spectra.Atmos. Chem.

Phys.,16, 10 195–10 213, doi:10.5194/acp-16-10195-2016.

Pommrich, R., and Coauthors, 2014: Tropical troposphere to stratosphere transport of carbon monoxide and long-lived trace species in the Chemical Lagrangian Model of the Stratosphere (CLaMS). Geosci. Model Dev.,7, 2895–2916, doi:10.5194/gmd-7-2895-2014.

Randel, W. J., M. Park, L. Emmons, D. Kinnison, P. Bernath, K. A.

Walker, C. Boone, and H. Pumphrey, 2010: Asian monsoon transport of pollution to the stratosphere. Science, 328, 611–613, doi:10.1126/science.1182274.

Schoeberl, M. R., L. C. Sparling, C. H. Jackman, and E. L. Fleming, 2000: A Lagrangian view of stratospheric trace gas distributions.

J. Geophys. Res.,105, 1537–1552, doi:10.1029/1999JD900787.

Stohl, A., 2006: Characteristics of atmospheric transport into the Arctic troposphere.J. Geophys. Res.,111, D11306, doi:10.1029/

2005JD006888.

Tissier, A.-S., and B. Legras, 2016: Convective sources of trajec- tories traversing the tropical tropopause layer.Atmos. Chem.

Phys.,16, 3383–3398, doi:10.5194/acp-16-3383-2016.

Vogel, B., and Coauthors, 2011: Transport pathways and signatures of mixing in the extratropical tropopause region derived from Lagrangian model simulations.J. Geophys. Res.,116, D05306, doi:10.1029/2010JD014876.

Waugh, D. W., and T. M. Hall, 2002: Age of stratospheric air:

Theory, observations, and models.Rev. Geophys.,40, 1–27, doi:10.1029/2000RG000101.

——, and Coauthors, 2013: Tropospheric SF6: Age of air from the Northern Hemisphere midlatitude surface.J. Geophys. Res.

Atmos.,118, 11 429–11 441, doi:10.1002/jgrd.50848.

Referenzen

ÄHNLICHE DOKUMENTE

In Bamako, Mali's capital, a brokered settlement between the military junta and an interim civilian government of the elites is not working.. Throughout the country, drought and

Es stellt sich daher nicht die Frage ob, sondern wie spezifische Positionierungen in Praktiken aktualisiert und transformiert werden, in denen sich die Mitglieder

The daily updating of the information on the web pages is a result of the updating of weather data and local information, which is done entirely by the individual countries..

The growing importance of teaching Industrial Art has also given great importance to learning through the observation and reproduction of the great national and international

Through the Oslo agreements the mainstream and majority Palestinian nationalists of Fatah recognised the state of Israel on 78% of Mandate Palestine, in a compromise that left

In order to the define the inverse, we define the alternating tableau of a multiset of necklaces M : associate with M the multiset of words which is the union with multiplicities of

On the basis of this model in the Institute of Economics and Industrial Engineering (Novosibirsk, USSR), the problem of bal- anced capital investment distribution

(E ff ective RF is de fi ned as the change in net top-of-the-atmosphere downward radiative fl ux after.. allowing for atmospheric temperatures, water vapor, and clouds to adjust, but