38
Contribution of Impervious Surfaces to Urban Evaporation P. Ramamurthy and E. Bou-Zeid Author Affiliations P. Ramamurthy and E. Bou-Zeid – Department of Civil and Environmental Engineering, Princeton University Primary Contact Prathap Ramamurthy E312 EQUAD, Dept. of CEE, Princeton University, Princeton, NJ 08544 Email- [email protected]

Contribution of Impervious Surfaces to Urban ... - Princeton

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Contribution of Impervious Surfaces to Urban ... - Princeton

Contribution of Impervious Surfaces to Urban Evaporation

P. Ramamurthy and E. Bou-Zeid

Author Affiliations

P. Ramamurthy and E. Bou-Zeid – Department of Civil and Environmental Engineering,

Princeton University

Primary Contact

Prathap Ramamurthy

E312 EQUAD, Dept. of CEE,

Princeton University,

Princeton, NJ 08544

Email- [email protected]

Page 2: Contribution of Impervious Surfaces to Urban ... - Princeton

Abstract

Observational data and the Princeton Urban Canopy Model, with its detailed

representation of urban heterogeneity and hydrological processes, are combined to study

evaporation and turbulent water vapor transport over urban areas. The analyses focus on

periods before and after precipitation events, at two sites in the Northeastern United

States. Our results indicate that while evaporation from concrete pavements, building

rooftops and asphalt surfaces is discontinuous and intermittent, overall these surfaces

accounted for nearly 18% of total latent heat fluxes (LE) during a relatively wet 10-day

period. More importantly, these evaporative fluxes have a significant impact on the urban

surface energy balance, particularly during the 48 hours following a rain event when

impervious evaporation is the highest. Thus, their accurate representation in urban

models is critical. Impervious evaporation after rainfall is also shown to correlate the

sources of heat and water at the earth surface, resulting in a conditional scalar transport

similarity over urban terrain following rain events.

1. Introduction

Evaporation has an overwhelming influence over local climatological, atmospheric and

agricultural processes; it has therefore been widely studied over the last century

[Brutsaert, 1982; Katul et al., 2012]. However, built surface evaporation over urban areas

remains one of the least studied and poorly understood topics in the fields of urban

hydrology and microclimatology. Oke [1982] commented on the role played by

impervious surfaces in influencing the surface energy fluxes nearly three decades ago,

Page 3: Contribution of Impervious Surfaces to Urban ... - Princeton

stressing that urban areas are not deserts; however, our current understanding of the

underlying processes driving urban evaporation and how they interact is still rudimentary.

Moreover, impervious surfaces are not the only built surfaces contributing to evaporation

in urban terrain; manmade vegetated surfaces or soils are also ubiquitous and their

hydrological and thermal properties can be quite distinct from those of surrounding

natural terrain.

The gaps in our understanding of the hydrological dynamics of these built surfaces can be

attributed to insufficient observations (compared to natural terrain) and the lack of

appropriate frameworks to interpret evaporative exchanges measured from the highly

heterogeneous footprints common in built terrain. This then results in a general scarcity

of thoroughly-tested small-scale models that resolve or represent these dynamics (we

review in the next section a few available models), and to inadequate or absent

representation of urban evaporation in many larger-scale weather and climate models.

This last point is a major area of concern as recent analyses have shown the role played

by urban areas in storm convergence and bifurcation [Bornstein and Lin, 2000; Dixon

and Mote, 2003; Shepherd, 2005] due to their high roughness and increased thermal

storage capacities, as well as the influence of urban surface representation on heavy

rainfall simulations in urban areas [Li et al., 2013]. Ignoring evaporation in such

simulations could erroneously strengthen the convergence leading to high precipitation

when the moisture availability is controlled by advection. In addition, evaporation is well

known to play a strong role in modulating the urban heat island, particularly during

extreme heat waves [Li and Bou-Zeid, 2013].

Page 4: Contribution of Impervious Surfaces to Urban ... - Princeton

2. Prior investigations and open questions

Eddy Covariance measurements have been commonly used to study evaporation from

urban areas [Grimmond et al., 2002; Rotach et al., 2005; Coutts et al., 2007; Vesala et al.,

2008; Hanna et al., 2011]. While these studies have helped us parameterize urban fluxes

and have advanced our general understanding of urban hydrology, the top-down approach

of EC systems invariably lumps the contributions from impervious surfaces (roofs, roads,

pavements, etc.) and vegetated surfaces (trees, bare soil, etc.) together; consequently, it

does not enable us to discern their individual contributions to urban evaporation.

Moreover, the inherent surface inhomogeneity often encountered in urban areas and the

resulting source/sink heterogeneity could lead to large scale advective errors [Grimmond

et al., 2002]; this is in addition to the challenges of interpreting EC data in the complex

urban flows [Aubinet et al., 2012]. These shortcomings of EC measurements have

restricted their capacity to yield a thorough understanding of the contributions of built

surfaces to urban evaporation.

On the modeling side, considerable advances have been made in representing evaporation

from urban areas, starting with the pioneering work of Grimmond and Oke [1986; 1986;

1991]. The urban evaporation-interception model described in Grimmond and Oke [1991]

is based on the Penman-Monteith-Rutter-Shuttleworth model [Penman, 1948; Monteith et

al., 1965; Rutter et al., 1972; Shuttleworth, 1978]. The urban surface in the model is

conceptualized as a single layer moisture store with parallel stores for pervious and

impervious surfaces. The model uses Penmann-Monteith equation to account for

evapotranspiration; the model also calculates drainage. The water balance for each surface

is then solved using the modified Rutter model. The Surface Urban Energy and Water

Page 5: Contribution of Impervious Surfaces to Urban ... - Princeton

Balance Scheme (SUEWS) developed by Jarvi et al., [2011] is also based on the urban

evaporation-interception model detailed above. The surface resistances in the SUEWS

model are parameterized explicitly for urban surfaces and the model uses an integrated

approach to account for urban vegetation. Other notable urban water balance models are

the Soil Model for Submesoscales, Urbanized Version 2 (SM2-U) [Dupont and Mestayer,

2006], and the Urban Hydrological Element (UHE) [Berthier et al., 2004] that expand the

range of represented hydrological processes.

Despite these efforts, the large majority of urban surface schemes that are currently adopted

in research or operational meteorological models continue to have a poor representation or

an absence of hydrological processes [Järvi et al., 2011]. Moreover, the general skill of

urban surface models, even those that represent hydrological processes relatively in detail,

to accurately model latent heat fluxes remains lacking compared to the other components of

the surface energy budget [Grimmond et al., 2010; 2011]. This indicates that major efforts

by the urban microclimatology and hydrology communities are still needed to develop a

more thorough understanding of urban hydrology and to effectively include this

understanding in models.

One class of urban models that is gaining increasing attention and that is hence useful to

target for improvement include the so-called urban canopy models (UCMs). UCMs, with

their single or multilayer representation of the urban surfaces, incorporate the effects of

urban morphology in modeling energy exchanges, and have been successfully used to

study urban energy budgets [Mills, 1993; Mills and Arnfield, 1993]. While UCMs have

been primarily employed in mesoscale models [Masson, 2000; Kusaka and Kimura,

2004] to study land-atmosphere interactions in urbanized areas, they have also been used

Page 6: Contribution of Impervious Surfaces to Urban ... - Princeton

in “offline mode” (i.e. stand-alone and driven by atmospheric observations) at smaller

spatial scales [Hamdi and Schayes, 2008; Bueno et al., 2013]. UCMs represent the street

canyon as an infinite rectangular cavity bordered by two buildings (usually of equal

height in the single layer representation; an infinite regular array of cubes can also be

represented with minor modifications to the model). They hence feature three facets: the

roofs, the walls, and the ground surfaces. Their geometric representation of urban terrain

allows them to accurately account for urban radiative exchanges. Many of the current

UCMs do account for latent heat fluxes from urban areas [Martilli et al., 2002; Lemonsu

et al., 2008; Ryu et al., 2011] and also for soil-freeze and thaw process [Leroyer et al.,

2010]. Evaporation surface fluxes in UCM are computed and coupled to the air aloft

using stability dependent aerodynamic resistances. The roof facet, which benefits from

maximum sky view factor at all periods, is directly coupled to the atmospheric reference

plane. The road (ground) and wall facets, impeded by buildings, are coupled to a

reference plane inside the canyon, which in turn is coupled to the reference plane above

the canyon. To account for vegetation, most UCMs use of one of two strategies: tiled or

integrated. In the tiling approach, a patch of vegetation, proportional to the urban

vegetated fraction, is allowed to evaporate independently. For example, in Lemonsu et

al.,[2007], a single layer UCM (Town Energy Budget, [Masson et al., 2002]) was

coupled to a land surface scheme (Interaction Soil-Biosphere-Atmosphere scheme

[Noilhan and Planton, 1989]) to account for evaporation from pervious surface. The

contribution from the vegetated patch is then fractionally weighted and added to the

overall surface energy budget. Another important example of the tiling approach can be

seen in the single or multilayer urban canopy representation implemented in the Weather

Page 7: Contribution of Impervious Surfaces to Urban ... - Princeton

Research and Forecasting model [Chen et al., 2011] where all urban hydrological processes

are simplistically represented as occurring over a patch of grass adjacent to the urban

canopy. The other strategy is to include the vegetation as part of the urban framework. In

this integrated approach, the pervious surface directly interacts with the built surfaces.

The urban vegetation model described in Lee and Park [2008] is one example that adapts

this strategy, which is much more realistic compared to the independent tile approach.

However, (1) as noted above, UCMs ability to accurately model latent heat remains

lacking compared to the other components of the surface energy budget [Grimmond et

al., 2010; 2011], and (2) the heterogeneity of the urban fabric at the sub-facet scale is not

represented and this omission can be one of the causes of the underperformance of these

models in view of the very striking variability in hygrothermal properties of urban

materials that makes the computation of effective averaged properties quite problematic.

These two points frame the scope of this paper, which aims to bridge the gap in our

understanding of hydrological processes in urban areas and to help improve urban

models. We apply a newly developed model that accounts for sub-facet heterogeneity and

that includes many hydrological processes (the Princeton Urban Canopy Model: PUCM)

over heterogeneous built terrain. We then compare it to eddy-covariance measurements to

investigate the challenges urban models face in representing latent heat fluxes. We

subsequently focus our attention on the influence of impervious surfaces (rooftops,

concrete pavements and asphalt roads and parking lots) on evaporative fluxes in urban

areas to understand their contribution to total evaporation and the importance of their

representation in models. The influence of hydrometeorological conditions on the urban

Bowen ratio (ratio of sensible to latent surface heat fluxes) and on the importance of

Page 8: Contribution of Impervious Surfaces to Urban ... - Princeton

impervious contribution to latent heat fluxes are then assessed, along with the

implications of these conditions for turbulent transport in the urban boundary layer and

what these implications suggest about the representation of turbulent fluxes in urban

models.

3. Modeling framework and experimental sites

PUCM is broadly based on the energy exchange scheme described by Kusaka et. al.,

[2001]. PUCM improves over conventional UCMs by its capability to distinguish various

urban materials (concrete, asphalt, grass, brick, gravel etc.) on each urban facet (ground,

roof and walls), and by the detailed hydrological module implemented to represent water

fluxes and storage [Wang et al., 2011a; 2013]. The improved PUCM continues to capture

complex processes such as radiative trapping, heat conduction in soil and impervious

surfaces, and turbulent exchanges of heat between the surface and the atmosphere as do

other UCMs. However, these exchanges are now computed for each sub-facet, consisting

of a different material, of a given urban facet. PUCM also models evaporation from soil

and vegetated surfaces by numerically solving the one-dimensional Richard’s equation

(see Ross and Parlange [1994] for an assessment of the numerical solution of this

equation) in a multi-layer soil matrix to obtain surface soil moisture, which is then used

to model evaporation using turbulent transfer functions and atmospheric specific

humidity (constrained by available energy). Potential evaporation is first computed and

then actual evaporation is reduced from its potential rate using a reduction factor based

on the moisture in the top layer for bare soils, while a stomatal resistance is added (to the

aerodynamic resistance represented by the turbulent transfer functions) for vegetation

[Brutsaert, 1982]. Impervious surfaces are assumed to have variable water storage,

Page 9: Contribution of Impervious Surfaces to Urban ... - Princeton

capped by a fixed upper limit capacity (bucket model); the storage is replenished by

precipitation and depleted by evaporation at a rate reduced from the potential evaporation

by a factor βI = current storage/maximum storage. The model, when used offline, is

forced by precipitation and hydrometeorological measurements (e.g. atmospheric specific

humidity). PUCM currently does not account for anthropogenic heat sources (although

they can be added as sources of heat inside or above the canyon if know), but does

account for dew formation when LE becomes negative, which has been reported as an

important term especially over rooftops [Richards, 2004]. Conduction in solids, including

soils, is modeled using a Green’s function based analytical solution of the 1D heat

equation. For additional information on the model, the reader is referred to Wang et al.

[2011a; 2013].

Apart from the model, data from an eddy-covariance (EC) station located inside the

campus of the University of Maryland Baltimore County (UMBC) is used for the

evaporation analysis (another station described later is used for turbulent transport

analyses). Figure 1 shows the observation site at UMBC (39.2542ºN 76.7097ºW) and the

footprint of the flux tower during convective conditions (when evaporation is maximum)

estimated using the approach of Hsieh et al. [2000]. To accurately match the modeled

results with the observations, the landcover within the footprint of the tower must be

represented as faithfully as possible in PUCM. The land surface is divided into four

sectors, each with distinct land cover characteristics; the model is run separately for each

sector; and the results are then aggregated into a single time series to compare to the

observations depending on the wind direction. The value of a variable in the aggregate

time series is equal to its value as predicted by PUCM over the upwind sector. These

Page 10: Contribution of Impervious Surfaces to Urban ... - Princeton

constructed time series are only used when comparing to observational data. When only

model output is being used, the outputs from the 4 sectors are averaged (weighted to

account for the different fractions covered by each sub-facet in different sectors) to get

values that represent the whole site. If we continue to use the time series based on wind

direction, the result will vary as the land use varies with wind direction and it will be

difficult to disentangle the effect of the hydrometeorology and diurnal forcing variability

from the effect of land use variation with wind. The averaging over all sectors is hence

needed to have results that are representative of the whole site and investigate the role

played by impervious surfaces towards evaporation in urban areas, before and after rainfall.

Table 1 details the fractions covered by each surface type as determined from aerial

photographs and the mean building height for each of the sectors; these parameters were

shown to be some of the most important in a sensitivity analysis of PUCM conducted

using Markov chain Monte-Carlo simulations by Wang et al. [2011b]. Roof albedo was

also shown to be an important input parameter in that study; for UMBC we determined a

representative albedo using aerial photographs and on-site inspections of multiple roofs

(measuring the albedos of all roofs was not practical). The black roofs (about 10% of

total roof area) were assigned an albedo of 0.1 and the light grey roofs (about 90% of

total roof area), which were found to be glossy/shiny, were assigned an albedo of 0.5.

These values are listed in table 2, which also details all the other parameters adopted in

the model runs. The water-holding capacity was also estimated by site inspection to be

about 1 mm for all ground impervious surfaces, and as 0.2 mm for roofs. The remaining

hygrothermal properties of the materials listed in the table were adopted from Wang et al.

[2011a; 2011b; 2013], who measured and calibrated for these properties for another site

Page 11: Contribution of Impervious Surfaces to Urban ... - Princeton

in the Northeastern United States. The other site is over the Princeton University Campus

and will later be used in this paper for turbulent transport analysis. Note here that the

roughness lengths assigned in PUCM are used to compute the exchange coefficients in

the model and hence are not directly comparable to values that one would measure for the

whole urban canopy using EC.

Figure 1: Map showing the land cover surrounding the UMBC flux tower and the sectoring scheme adopted

in the PUCM (reproduced with permission from the American Meteorological Society from Ramamurthy et al. [2014]).

Table 1: Land cover characteristics of the four sectors.

Land Cover Characteristics

Sector 1 Roof 25%, Asphalt 32%, Concrete13%, Green 30%

Mean Building Height 11 m

Sector 2 Roof 1%, Asphalt 40%, Concrete10%, Green 49%

Mean Building Height 6 m

Sector 3 Roof 40% (12% black, 28% grey), Asphalt 10%, Concrete 29%, Green 21%

Mean Building Height 12 m

Sector 4 Roof 33% (7% black, 28% grey), Asphalt 9%, Concrete 24%, Green 34%

Mean Building Height 8 m

Page 12: Contribution of Impervious Surfaces to Urban ... - Princeton
Page 13: Contribution of Impervious Surfaces to Urban ... - Princeton

Table 2: Input parameters used in the model for all sectors

Properties Values

Roof surface albedo 0.50 for grey, 0.1 for black Thermal roughness length for roof surface 0.001 m Thermal roughness length for canyon-to-air exchanges 0.005 m Momentum roughness length for roof surface 0.01 m

Momentum roughness length for canyon-to-air surfaces 0.05 m Water-holding capacity for ground concrete and asphalt pavements 0.001 m

Water-holding capacity for roofs 0.0002 m

Wall surface albedo [brick] 0.25 Ground surface albedo [asphalt; concrete; grass] [0.15; 0.40; 0.10] Roof surface emissivity 0.95 Wall surface emissivity [brick] 0.95 Ground surface emissivity [asphalt; concrete; grass] [0.95; 0.95; 0.93] Roof heat capacity 1.8×106 JK–1m–3 Wall heat capacity (brick) 1.2×106 JK–1m–3

Ground heat capacity [asphalt; concrete; green] [1.0; 2.4; 1.2]×106 JK–1m–3

Roof thermal conductivity 0.6 WK–1m–1

Wall thermal conductivity 1.3 WK–1m–1 Ground thermal conductivity [asphalt; concrete; green] [1.2; 1.8; 1.2] WK–1m–1 Internal building temperature 23 ºC Initial wall temperature 19 ºC Initial ground surface temperature 21 ºC Initial volumetric soil moisture 0.2 m3 m–3 Saturated soil water content 0.33 m3 m–3 Residual soil water content 0.06 m3 m–3 Saturated soil hydraulic conductivity 3.38×10–6 m s–1

PUCM also needs basic forcing meteorological variables such as incoming radiation,

temperature, wind speed, specific humidity and precipitation, which were measured at the

flux tower using the sensors listed in table 3. In addition to the instruments tabulated in

that table, a Campbell Scientific CSAT3 sonic anemometer was employed in conjunction

with a LICOR LI7500 open path gas analyzer to quantify latent and sensible heat. To

calculate the fluxes, standard EC post-processing techniques [Foken et al., 2005] were

Page 14: Contribution of Impervious Surfaces to Urban ... - Princeton

followed. The instantaneous fields were corrected for spikes by manual inspection and

periods contaminated by precipitation were removed (using the Automatic Gain Control

(AGC) filter, standard on the LI7500). The time series were then linearly detrended and a

planar fit procedure described by Wilzack et al. [2001] followed to rotate the coordinates

of the wind field. The Reynolds-averages, the concomitant perturbations, and the fluxes

were computed over 30-minute periods. The WPL corrections [Webb et al., 1980] were

finally applied to the mean fluxes to account for errors related to density fluctuations.

PUCM runs yielded data at 5-minute intervals. The data were then averaged into 30-

minute bins to compare to the eddy covariance observations. To obtain a monthly-

averaged diurnal cycle, both modeled and observed fluxes were separated by day, and the

corresponding half-hour bins from each day were ensemble-averaged.

Table 3: Meteorological data used to force PUCM at the UMBC site

Meteorological Fields Instrumentation Incoming longwave (LW) &

shortwave (SW) radiation1 Kipp & Zonen CNR1 (Kipp & Zonen Inc., Delft, Netherland),

which includes the CM3 thermopile pyranometer and the CG3

pyrgeometer. Air Temperature Vaisala HMP 45 (Vaisala Inc., Vantaa, Finland) Specific Humidity Vaisala HMP45 (Vaisala Inc., Vantaa, Finland) Wind Speed Campbell Scientific CSAT3 (Campbell Sci. Inc., Logan, UT,

USA) Atmospheric Pressure Licor 7500 (Licor Inc., Lincoln, NE, USA) Precipitation Raingauge 6011-A (All Weather Inc., Sacremento, CA, USA)1

These incoming radiations measurements were made at the meteorological tower at Cub Hill, 20 km from UMBC [Crawford et al., 2011].

4. Model Validation

Figure 2 compares the modeled latent and sensible heat flux (their mean diurnal cycles)

to the EC observations for the dry period from July 1-20 (a) and the wet period from July

Page 15: Contribution of Impervious Surfaces to Urban ... - Princeton

21-31 (b). During the dry period, the model captures the sensible heat flux well in the

morning but has biases in the afternoon; on the other hand, the latent heat flux is slightly

overpredicted in the morning and underpredicted in the afternoon. During the wet period,

the observational data are more scattered, but the model seems to predict the fluxes well,

albeit with some overprediction of latent heat fluxes. Quantitatively, the root mean square

errors (RMSEs) comparing the modeled and observed fluxes for the dry period were 33

Wm–2 for H and 23.5 Wm–2 for LE. For the wet period, the RMSEs were 24.1 Wm–2 for

H and 36.4 Wm–2 For LE. The Mean Bias Errors (MBEs) in H and LE for the dry period

were 1.1 Wm–2 and 1.6 Wm–2, respectively, and for the wet period the MBEs were 4

Wm–2 for H and 2 Wm–2 for LE indicating very slight underestimation of the daily-

averaged H and LE. These values are comparable to the ones obtained by other UCMs

with realistic hydrological representations [Ryu et al., 2011].

Some observational aspects that could affect this comparison include the fact that the

radiation was not measured directly at the site but rather about 20 km away at the Cub

Hill tower; temporal shifts in radiative patterns (particularly under a patchy cloud cover)

can have an impact on the model results. In addition, for the last 10 days with significant

precipitation, the periods with no measurements and the small sample size (less than 10

data points per 30-minute period) result in scatter in the averaged fluxes that would be

smoothed out with a larger sample. These observed biases might also be related to model

limitations that pertain to the turbulent transfer function parameterizations in the model

and how they depend on stability, to the thermal properties used for the materials, or to

the simplicity of the infinite 2D canyon geometric representation of the complex real-

world canopy. Another aspect of the model that might contribute to the discrepancy is its

Page 16: Contribution of Impervious Surfaces to Urban ... - Princeton

inability to account for transpiration from deep-rooted plants. Given the heterogeneity of

the urban surfaces and the complexity of transpiration process for various tree and plant

species [Brutsaert, 1982; 2005], these effects might be hard to replicate in the model

where vegetated surfaces are assumed to consist exclusively of grass (although work to

better represent the various vegetation types is underway).

Figure 2: Comparison of mean diurnal cycles of the modeled latent (LE) and sensible (H) heat fluxes to

flux tower observations. Plot in (a) is for the dry period (July 1-20, 2009) and in (b) is for the wet period

(July 21-31 2013).

However, despite the limitations of the model in representing the urban

evapotranspirative exchanges with very high accuracy and despite the uncertainty of the

input parameters, some positive aspects of the comparison are worth noting: the MBE’s

are quite small indicating that the model can represent the daily-averaged fluxes quite

well; the change in the Bowen ratio between the dry and wet periods is very-well

captured by PUCM; the positive values of LE and the negative values of H during

nighttime are reproduced by the model; and the peak flux times during the wet period are

captured accurately.

Page 17: Contribution of Impervious Surfaces to Urban ... - Princeton

5. Impact of built surfaces on urban evaporation

In the previous section, PUCM was set up and evaluated for the ensemble-averaged

diurnal cycles for wet and dry days. To further assess the model performance and to

illustrate the response of built surfaces to precipitation, the last nine days of the month,

which featured several rain events, are chosen for further analysis. The observed and

simulated fluxes for these days are shown in figure 3. The model captures the sensible

heat flux quite well; its performance in reproducing the latent component of the flux is

less satisfactory, but remains acceptable. The RMSE values for H and LE during this

period were 23 Wm–2 and 38 Wm–2, respectively. Notice also that rainfall was not

recorded during all data gaps periods, but data gaps strongly indicate rainfall in this

period since they occur when the LI7500 windows are wet. Hence, this discrepancy

suggests that the used tipping bucket rain gage, with a sensitivity of 0.25 mm per tip, may

have missed some small rain events and hence provided a biased input to PUCM. The

first precipitation event occurs around midday on July 23. This event results in low H and

very high LE after rainfall ends and the instruments dry and resume their normal

functioning. PUCM reproduces this increase in LE but the model outputs are smoothed

relative to the observations. Similarly, the precipitation event on the night of the 24th

results in depressed H and in intermittent LE, which is higher than H, during the next day.

The very high values of LE in the observations after the data gap periods are a feature

worth elaborating on. Observational data gaps correspond to rain periods and the peaks

after such periods clearly indicate intensive evaporation from the impervious surfaces

that have stored heat and are able to supply this heat back to sustain high evaporation

rates after a rain shower. Tests not shown here further indicate that if the evaporation

Page 18: Contribution of Impervious Surfaces to Urban ... - Princeton

from the impervious fraction is not accounted for, PUCM produces higher H and lower

LE, and the match with the observational data is noticeably inferior to the one depicted in

the figure. The combination of these observational and simulation results supports the

hypothesis that impervious surface evaporation is significant in urban terrain. Though

some features of the model need further development and some input parameters need to

be carefully and accurately determined for each site, overall, the performance of PUCM

suggests that it captures the physics of the urban energy and water budgets with sufficient

realism to enable its use as a framework to investigate the impact of various urban

parameters on evaporation, as will be done next.

Figure 4 shows the contribution of various built surfaces to overall evaporation from

PUCM simulations for the last nine days of the study period. Note that from this point on

we will not use model output aggregated by sector to mimic observation data, but rather

the outputs averaged over all sectors that will better represent the urban surface dynamics

and will have invariable land use. But it is still informative to note that the peak fluxes

from the vegetated areas are higher than the area-averaged fluxes seen in Figure 3, while

the peak fluxes from the impervious sub-facets tend to be lower. The bars in the figure

indicate the magnitude and the period of rain events. During the drier days (July 21st, 27th

and 28th), LE from the built surfaces was zero. The water retained on the impervious

surfaces had completely evaporated and the vegetated surfaces are responsible for all of

the latent heat flux. However, the roofs at other urban sites (e.g. the Princeton site we

discuss later in this paper) and some other urban materials can have significantly higher

water-holding capacities and can hence sustain evaporation for much longer. Gravel

Page 19: Contribution of Impervious Surfaces to Urban ... - Princeton

roofs, which are very common in the Northeastern US, could have water-holding

capacities of up to 10 mm due to dirt accumulation and pore clogging.

Figure 3: Modeled and observed latent (upper panel) and sensible (lower panel) heat fluxes at the UMBC

site for a 9-day period during July 2009. The x-axis is the local time/date and the date tick marks

correspond to 00:00 hours. The gaps observed in measured H and LE are due to the EC systems inability to

monitor fluxes during and immediately following precipitation events. The rain events particularly affect

the Licor gas analyzer.

Page 20: Contribution of Impervious Surfaces to Urban ... - Princeton

Figure 4: Modeled latent heat fluxes from different surfaces at the UMBC site for July 2009. The x-axis is

the local time/date and the date tick marks correspond to 00:00 hours.

During wet days, when the impervious surfaces are covered with water, their LE fluxes

are a significant fraction of the total evaporation as shown in Figure 4. LE from asphalt

and roof surfaces peak rapidly and almost shut off at night, while LE from concrete

surfaces persists for a while, with a lag, and is sustained at night by the high thermal

energy stored in the concrete (which has a very high thermal effusivity and hence has

more energy stored that can be later release for evaporation). These trends are a result of

the different thermal properties of the materials which result in different surface

temperatures and consequently in different saturated specific humidity over the wet

impervious surfaces. Figure 5 illustrates this by comparing the surface temperature of

different sub-facets for July 23-31. It is clear from the plot that while asphalt experiences

the highest daytime temperatures leading to the highest LE and a rapid depletion of its

water storage capacity, concrete maintains a lower temperature during daytime and a

higher temperature during nighttime due to its high thermal effusivity resulting in a

broader evaporation cycle. Notice the rapid synchronized cooling of the modeled surface

Page 21: Contribution of Impervious Surfaces to Urban ... - Princeton

temperatures and observed air temperature after the large second precipitation event, as

well as the fact that the surface temperatures of concrete and asphalt remain higher than

that of air all through the night.

Figure 5: Modeled surface temperature from different surfaces at the UMBC site for July 2009. The x-axis

is the local time/date and the date tick marks correspond to 00:00 hours.

These results also illustrate the highly intermittent fluxes observed for the built surfaces,

which are in contrast to the more continuous green cover evaporation. This is due to the

fact that while the green cover evaporation is moderated by energy and water infiltration

and subsequent release from the soil, the built surfaces evaporate at potential rate

following rain events and have lower water storage capacity. Overall LE from roofs is

considerably lower than from concrete or asphalt pavements due to the lower water-

holding capacity assigned to roofs (0.2 mm versus 1 mm for ground impervious sub-

facets). But even when all surfaces are wet (water-holding capacity is not relevant then),

roof evaporation rates are lower compared to the rates from concrete and asphalt surface,

despite the maximum sky view factor and lower total transport resistance of the roofs

30 31 0115

30

45

60

July 2009

° C

0

3

6

9

12

mm

TAirVegetationAsphaltConcreteRoofRain

Page 22: Contribution of Impervious Surfaces to Urban ... - Princeton

compared to the shaded (from the sun and wind) asphalt and concrete ground surfaces.

This is explained by the roof’s lower temperature throughout the day (figure 5), which is

directly attributable to the high albedo of the white roofs that constitute 90% of total roof

area at UMBC.

It is also worth here discussing the implications of the findings on EC flux measurements

over any terrain. The observed flux data plotted in Figure 3 has quite a few gaps, in fact

nearly 45 hours of data were lost due to rain during the second half of July 2009. During

these periods, when no LE was recorded by the EC system, high LE periods are visible in

the modeled data. Traditional EC techniques fail in accounting for LE immediately

following rain events [Heusinkveld et al., 2008]. During and immediately after rain, the

open path sensor LI7500 is not reliable since its glass windows get saturated and it takes

30-60 minutes, depending on the surrounding weather, for the windows to dry and return

to normal operation. Sonic anemometers also cannot produce reliable measurements

under heavy rain, their combination with closed-path gas analyzers (which are not

affected by rain) for flux measurements during rainy periods is also going to miss some

important flux periods. During these periods, as observed in our modeling data,

significant evaporation happens, with a significant contribution from impervious

surfaces, and goes unaccounted for with the traditional measurement techniques. For the

UMBC simulation, we removed the modeled LE data during periods that exceeded the

safe AGC limits and found that nearly 5% of the monthly evaporation occurred during

those periods immediately following the rain events. This number is significant and

cannot be discounted; similar effects can be expected for CO2 fluxes after rain. Hence,

carbon flux networks might suffer biases due to the omission of post-rain periods.

Page 23: Contribution of Impervious Surfaces to Urban ... - Princeton

6. Influence of hydrometeorological factors

Urban areas, in general are assumed to alter the Bowen ratio (β = H/LE) by increasing the

contribution of the sensible heat flux relative to the latent heat flux [Christen and Vogt,

2004; Offerle et al., 2006; Coutts et al., 2007]. While this is true on most occasions,

under wet conditions one anticipates that the urban Bowen ratio will be reduced due to

the stronger evaporation from the impervious parts. July 2009 in Baltimore was an

interesting month to study since the first half of the month was dry, while, in stark

contrast, the second half experienced extended precipitation periods. Following a 14 mm

shower on July 1, 2009, the next twenty days were just dry. On the other hand, the final

ten days had nearly 40 mm of total rainfall.

Figure 6 shows the mean diurnal cycle of the Bowen ratio (with 30 minute periods) for

the dry and wet periods predicted by the model for the whole modeled urban surface. The

overall average for the vegetated surface is also shown as a reference. The daytime ratio

averaged around 0.5-0.75 for the wet period and over 1 for the dry period. The results

observed underline the influence of precipitation on β, and the cooling role of impervious

evaporation in urban terrain during wet periods. The figure also shows that over

vegetated surfaces β remains below 0.5 over the entire day.

Page 24: Contribution of Impervious Surfaces to Urban ... - Princeton

Figure 6: Modeled 30-min ensemble-averaged daytime Bowen ratio for wet (July 21-31, 2009) and dry

(July 1-20, 2009) periods at UMBC for July 2009.

7. Impact of impervious evaporation and hydrometeorological conditions on

turbulent transport in the urban boundary layer

The previous sections applied PUCM and analyzed EC data to understand how the

physical properties of the built surfaces and the hydro-meteorological conditions

influence evaporation in urban areas. An implicit assumption in the modeling part, as in

all UCMs, was that atmospheric turbulence is equally efficient in producing sensible and

latent heat exchanges between the surface and the ABL, but many studies show that this

is far from being a generally applicable assumption (see the review of the seminal work

of Wilfried Brutsaert and collaborators on this topic in the review paper in this special

issue by Dias [2013]). This turbulent transfer efficiency affects the exchange coefficients

(turbulence transfer functions or aerodynamic resistances) in PUCM and other urban

models used to quantify surface fluxes. Given that the sum of the latent and sensible heat

fluxes is constrained by the available energy, the ratio of sensible to latent heat exchange

coefficient and the associated ratio of sensible to latent heat transfer efficiency are very

Page 25: Contribution of Impervious Surfaces to Urban ... - Princeton

critical for correctly partitioning surface fluxes between latent and sensible heat (if the

ratio is correct, the model will perform relatively well even if the magnitude of the two

coefficients have larger errors than their ratio). This final section is thus dedicated to the

investigation of this scalar transport similarity over urban terrain, and in particular how it

is influenced by precipitation events that activate evaporation from impervious surfaces.

One feature of urban fabric that is highly susceptible of causing a breakdown of scalar

similarity is the multiplicity and heterogeneity of sources and sinks for heat and water

vapor [Roth and Oke, 1995; Asanuma and Brutsaert, 1999; Nordbo et al., 2013]. To

investigate this similarity, or lack thereof, we use EC data from a grass covered and a

densely built site, less than 1 km apart in New Jersey (NJ). Figure 7 shows an aerial view

of both the sites. The UMBC site is not as urbanized as the NJ built site and we had no

data over a nearby natural homogeneous surface to conduct the analyses we planned

using UMBC data, hence the use of these other sites. The first is, the Princeton campus

EC station located in a highly dense urban environment surrounded by built surfaces

(40.3509°N, –74.6510°W). The measurements were made at 23.5 m above ground level

(AGL). The roof was 18 m AGL (surrounding building were lower) and the LICOR 7500

and the Campbell Scientific CSAT3 anemometers were mounted on a 5 m tall triangular

lattice tower fixed on an elevated protrusion of the roof. While the authors are aware of

the flow distortion effects that particularly influence measurements made over rooftops

[Aubinet et al., 2012] when the sensors are not fully and unambiguously above the

roughness sublayer, the relatively low placement of these sensors was mainly aimed at

keeping the footprint small and highly built. Moreover flow distortion effects are not

critical to the analyses we will perform here, as illustrated later, since for these sites we

Page 26: Contribution of Impervious Surfaces to Urban ... - Princeton

are not concerned with accurately computing fluxes. The Broadmead tower (40.3463°N,

–74.6434°W), 3m AGL, is located in much more homogenous terrain surrounded by a

grass field (with an average roughness height of 0.1 m). Both these sites used a LICOR

7500 and Campbell Scientific CSAT3 anemometers to measure the instantaneous

velocity and scalar fields. Standard EC post-processing techniques exactly similar to the

ones detailed in the methods section for the UMBC site were followed to compute the

fluxes. While the measured average Bowen Ratio at the Broadmead site was around 0.4

for the summer months during the convective period (z/L < –0.1, where z is the height at

which the measurements were made and L is the Monin-Obukhov lengthscale), the

Bowen Ratio at Princeton was more than 2, clearly illustrating the contrast in the surface

fluxes at the two sites.

Figure 7: Aerial photo depicting the locations of the two flux towers and the surrounding land cover

characteristics. The picture on the left (a) shows the Broadmead site and the one on the right (b) shows the

Princeton site.

Figure 8 compares the transfer efficiencies of heat and water vapor, expressed as the

correlation functions (see further analyses and interpretations of these efficiencies in Roth

and Oke [1995] and Li and Bou-Zeid [2011])

RwT =′w ′T

σwσT, (2)

Page 27: Contribution of Impervious Surfaces to Urban ... - Princeton

Rwq =′w ′q

σwσ q

(3)

at the homogenous Broadmead site. Here, w′, T′ and q′ represent the fluctuations from the

means of vertical velocity, temperature and specific humidity, respectively, and the σ’s

indicate the standard deviations. For the period of the comparison, the Rwq and RwT values

at Broadmead were identical during the daytime and this was generally true all through

the growing season at the Broadmead site. The near identical values indicate a high

degree of correlation, resulting from high correlation of the surface sources of the two

scalars and yielding a high degree of similarity in the turbulence transfer of heat and

water vapor at the Broadmead site.

Figure 9 compares RwT and Rwq for the dense urban site located inside the Princeton

university campus. During the dry periods, July 5-8, a distinct difference in the transfer

efficiencies is visible, with a higher transport efficiency for sensible heat compared to

water vapor. This dissimilarity is in agreement with the comparison of the two

efficiencies over built areas by Roth and Oke [1995] who underlined that these trends

observed over urban terrain are different from the ones found over vegetated terrain; they

reported that the source-sink dissimilarities at the surface kept the ratio of RwT / Rwq

always above 1, irrespective of atmospheric stability. Buoyancy effects in the flow

enhance the transfer of sensible heat [Li et al., 2012]; when the sources of the two scalars

are well correlated at the surface, water vapor benefits from this correlation and its

transfer efficiency is enhanced as it rides along with sensible heat in the buoyant plumes

that develop due to surface sensible heat flux. But this similarity breaks down when the

surface sources are decorrelated as happens over the urban areas.

Page 28: Contribution of Impervious Surfaces to Urban ... - Princeton

However, when forced by rain, as seen on July 9th, RTq for the Princeton flux site

approaches unity and becomes very similar to the value at the Broadmead site. This

observed similarity is a consequence of the water retaining capacity of the roof surfaces

and other manmade impervious surfaces at the site. During the dry periods the built

surfaces are strictly sources of heat and the LE observed during this period is mainly

derived from the surrounding green cover. However, during wet periods, the

impermeable surfaces that dominate the Princeton site evaporate, allowing higher spatial

correlation of the surface fluxes and resulting in a conditional scalar transport similarity

(conditional in the sense that it derives from the surface correlation after rain of the two

scalars rather than from dynamical considerations). Results not shown here also clearly

demonstrate that the correlation of the instantaneous sensible heat and water vapor fluxes

(this is the correlation coefficient of the instantaneous products w′T′ and w′q′) also

increases from a value of about 0.2 to a value exceeding 0.8 after rain. This correlation

coefficient is a measure of whether the eddies transporting sensible heat are the same as

the ones transporting water vapor, and its rise indicates that these eddies are distinct

under dry conditions but strongly overlapping under wet conditions.

Page 29: Contribution of Impervious Surfaces to Urban ... - Princeton

Figure 8: 30-minute averaged transfer efficiencies of heat and water vapor at the Broadmead site.

Figure 9: 30-minute averaged transfer efficiencies of heat and water vapor at the Princeton site. The dotted

box highlights similar transfer efficiencies for heat and water vapor following precipitation.

These results suggest that the Bowen Ratio (β) at the urban site influences the transport

efficiencies quite significantly and could be used as a parameter to adjust the relative

values of the heat and water vapor transfer functions in UCMs (since it is predicted by

UCMs). Figure 10 depicts the variations of RwT and Rwq with β at the Princeton site (for

Rw

x

0

1

2

3

4July 2011

Rai

n (

mm

)

RainR

wt

Rwq

Page 30: Contribution of Impervious Surfaces to Urban ... - Princeton

the convectively active period only, around 51% of the overall data) and data from the

entire month of July 2011 were utilized. The plot shows that for lower β values, between

0.1 and 2, the RwT and Rwq are somewhat similar; however, as the β value increases, a

severe drop in Rwq is observed. In fact, for β values greater than 2, the Rwq values drop

below 0.2. The RwT values on the other hand remain steady or rise slightly with increasing

β values. These findings agree with the trends reported by Wang et al. [2013], where

analyses of turbulence data collected over central Beijing from a 300 m tall tower also

indicate a higher transport efficiency for sensible heat under convective conditions. Their

results further indicate that when the atmosphere is very mildly unstable (which could be

related to periods following rain events), the ratio of the efficiencies is closer to one and

water vapor transport could in fact become more efficient that sensible heat transport

under neutral conditions. The findings of this and the other studies referenced above in

general confirm that temperature is the main active scalar modulating the turbulence

(from its baseline shear-generated state under neutral conditions) and changing its

transport characteristics under unstable conditions. Water vapor simply rides with

temperature when the two scalars are correlated at the surface, while its transport is

hampered when it is not well correlated with temperature, as also concluded by Li et al.

[2012] for heterogeneous natural surfaces.

Page 31: Contribution of Impervious Surfaces to Urban ... - Princeton

Figure 10: Variation of RwT and Rwq with Bowen Ratio at the Princeton site for July 2011.

7. Conclusions

Hydrological processes develop and interact in complex and non-linear ways, but they

are vital to a broad range of human activities and interests. Wilfried Brutsaert and Jean-

Yves Parlange, to which the special issue where this paper appears is dedicated, devoted

their productive academic careers largely to the advancement of our understanding of

these processes and laid the groundwork for hydrological sciences to meet the challenges

of the 21st century. Prominent among these challenges is the interaction of manmade and

natural systems in urban terrain. The strong modification in hygrothermal properties and

the striking increase in surface variability are two features of the urban fabric that make it

particularly challenging from a hydrological modeling perspective. One here has to

underline the pioneering role of Wilfried Brutsaert in elucidating the complexities of

land-atmosphere interaction over spatially-variable terrain.

In this study, we illustrate some of these urban hydrological complexities using the newly

developed Princeton Urban Canopy Model (PUCM), in conjunction with observed flux

Rwx

Page 32: Contribution of Impervious Surfaces to Urban ... - Princeton

data to probe the skill of urban models in reproducing urban evaporation and to advance

our understanding of urban hydrology. The model’s performance was quite appropriate to

justify its use for further probing of the role of impervious surfaces in producing

evaporative fluxes from built terrain: mean bias errors averaged over multiple days varied

between wet and dry periods and for sensible or latent fluxes, but remained under 4 W

m-2. However, the uncertainty in the urban physical properties remains a challenge for

urban modeling.

While impervious surfaces have been often characterized as having a high Bowen ratio

due to their thermal storage capacity and relative inability to store water, our results show

that when forced by precipitation, their water retention ability, though limited, can sustain

significant evaporation (a simple computation can show that a 1 mm water film can

sustain a latent heat flux of about 85 Wm–2 for 8 hours as it evaporates completely).

Impervious surfaces are able to evaporate at higher rates than vegetated surface when wet

due to their higher storage of heat prior to rain events; also we noted that some

impervious surfaces such as older gravel roofs can have water storage capacities of about

10 mm. Evaporation from different imperious facets displayed distinct temporal patterns

related to the distinct thermal properties of urban materials. Overall, the variability of the

thermal properties results in temporally broader daytime evaporative and energetic

cycles. This impervious evaporation was shown to significantly alter the Bowen ratio in

urban areas during wet periods. The results show that during a wet 10-day period nearly

17% of total LE was from built surfaces, although they only evaporated following

precipitation events.

Page 33: Contribution of Impervious Surfaces to Urban ... - Princeton

We then proceeded to investigate the similarity of scalar (temperature and water vapor)

transport in urban terrain during dry and wet periods. Since PUCM and all urban models

the authors are aware of assume that these two scalars are transported with equal

efficiency, examining this assumption is very important. While urban areas have been

known in general to break the similarity in scalar transport due to the strong variability in

surface sources and sinks of sensible and latent heat, our analysis shows that scalar

transport similarity is recovered during wet periods due to the increased correlation of

surface sources and sinks for the two scalars. But one has to note that urban models have

to address and account for this dissimilarity during dryer periods when the turbulent

sensible heat transport is much more efficient than water vapor transport.

Acknowledgement

This work is supported by NSF under Grant No. CBET-1058027. We would like to thank

Professor Claire Welty and her research team at the University of Maryland at Baltimore

County for sharing the UMBC data, and Professor Jim Smith and Dr. Mary Lynn Baeck

from Princeton University for sharing the Broadmead data.

References

Asanuma, J., and W. Brutsaert (1999), The effect of chessboard variability of the surface fluxes on the aggregated turbulence fields in a convective atmospheric surface layer, Boundary-Layer Meteorology, 91(1), 37–50.

Aubinet, M., T. Vesala, and D. Papale (2012), Eddy covariance: a practical guide to measurement and data analysis, Springer.

Berthier, E., H. Andrieu, and J. D. Creutin (2004), The role of soil in the generation of urban runoff: development and evaluation of a 2D model, Journal of Hydrology, 299(3-4), 252–266.

Bornstein, R., and Q. Lin (2000), Urban heat islands and summertime convective thunderstorms in Atlanta: three case studies, Atmospheric Environment, 34(3), 507–

Page 34: Contribution of Impervious Surfaces to Urban ... - Princeton

516.

Brutsaert, W. (1982), Evaporation into the atmosphere: theory, history, and applications, Reidel Dordrecht.

Brutsaert, W. (2005), Hydrology: An Introduction, Cambridge University Press.

Bueno, B., J. Hidalgo, G. Pigeon, L. Norford, and V. Masson (2013), Calculation of Air Temperatures above the Urban Canopy Layer from Measurements at a Rural Operational Weather Station, Journal of Applied Meteorology and Climatology, 52(2), 472–483.

Chen, F., H. Kusaka, R. Bornstein, J. Ching, C. Grimmond, S. Grossman Clarke, T. Loridan, K. W. Manning, A. Martilli, and S. Miao (2011), The integrated WRF/urban modelling system: development, evaluation, and applications to urban environmental problems, International Journal of Climatology, 31(2), 273–288.

Christen, A., and R. Vogt (2004), Energy and radiation balance of a central European city, International Journal of Climatology, 24(11), 1395–1421.

Coutts, A. M., J. Beringer, and N. J. Tapper (2007), Impact of Increasing Urban Density on Local Climate: Spatial and Temporal Variations in the Surface Energy Balance in Melbourne, Australia, Journal of Applied Meteorology and Climatology, 46(4), 477–493.

Crawford, B., C. S. B. Grimmond, and A. Christen (2011), Five years of carbon dioxide fluxes measurements in a highly vegetated suburban area, Atmospheric Environment, 45(4), 896–905.

Dias, N. (2013), Research on atmospheric turbulence by Wilfried Brutsaert and collaborators, Water Resources Research, 49 (11), 7169–7184.

Dixon, P. G., and T. L. Mote (2003), Patterns and causes of Atlanta's urban heat island-initiated precipitation, Journal of Applied Meteorology, 42(9), 1273–1284.

Dupont, S., and P. G. Mestayer (2006), Parameterization of the Urban Energy Budget with the Submesoscale Soil Model, Journal of Applied Meteorology and Climatology, 45(12), 1744–1765.

Foken, T., M. Göockede, M. Mauder, and L. Mahrt (2005), Post-Field Data Quality Control, in Handbook of Micormeteorology, pp. 181–208, Kluwer Academic Publishers.

Grimmond, C. S. B., and T. R. Oke (1986), Urban Water Balance: 2. Results From a Suburb of Vancouver, British Columbia, Water Resources Research, 22(10), 1404–1412.

Grimmond, C. S. B., and T. R. Oke (1991), An evapotranspiration-interception model for

Page 35: Contribution of Impervious Surfaces to Urban ... - Princeton

urban areas, Water Resources Research, 27(7), 1739–1755.

Grimmond, C. S. B., T. R. Oke, and D. G. Steyn (1986), Urban Water Balance: 1. A Model for Daily Totals, Water Resources Research, 22(10), 1397–1403.

Grimmond, C. S. B., T. S. King, F. D. Cropley, D. J. Nowak, and C. Souch (2002), Local-scale fluxes of carbon dioxide in urban environments: methodological challenges and results from Chicago, Environmental Pollution, 116(1), S243–S254.

Grimmond, C., M. Blackett, M. Best, J. Barlow, J. Baik, S. Belcher, S. Bohnenstengel, I. Calmet, F. Chen, and A. Dandou (2010), The international urban energy balance models comparison project: first results from phase 1, Journal of Applied Meteorology and Climatology, 49(6), 1268–1292.

Grimmond, C., M. Blackett, M. J. Best, J. J. Baik, S. E. Belcher, J. Beringer, S. I. Bohnenstengel, I. Calmet, F. Chen, and A. Coutts (2011), Initial results from Phase 2 of the international urban energy balance model comparison, International Journal of Climatology, 31(2), 244–272.

Hamdi, R., and G. Schayes (2008), Sensitivity study of the urban heat island intensity to urban characteristics, International Journal of Climatology, 28(7), 973–982.

Hanna, S., E. Marciotto, and R. Britter (2011), Urban energy fluxes in built-up downtown areas and variations across the urban area, for use in dispersion models, Journal of Applied Meteorology and Climatology, 50(6), 1341–1353.

Heusinkveld, B. G., A. F. G. Jacobs, and A. A. M. Holtslag (2008), Effect of open-path gas analyzer wetness on eddy covariance flux measurements: A proposed solution, Agricultural and Forest Meteorology, 148(10), 1563–1573.

Hsieh, C., G. Katul, and T.-W. Chi (2000), An approximate analytical model for footprint estimation of scalar fluxes in thermally stratified atmospheric flows, Advances in Water Resources, 23(7), 765–772.

Järvi, L., C. S. B. Grimmond, and A. Christen (2011), The Surface Urban Energy and Water Balance Scheme (SUEWS): Evaluation in Los Angeles and Vancouver, Journal of Hydrology, 411(3-4), 219–237.

Katul, G. G., R. Oren, S. Manzoni, C. Higgins, and M.B. Parlange (2012), Evapotranspiration: A process driving mass transport and energy exchange in the soil-plant-atmosphere-climate system, Review of Geophysics, 50(3), RG3002.

Kusaka, H., and F. Kimura (2004), Coupling a Single-Layer Urban Canopy Model with a Simple Atmospheric Model: Impact on Urban Heat Island Simulation for an Idealized Case, JMSJ, 82(1), 67–80.

Kusaka, H., H. Kondo, Y. Kikegawa, and F. Kimura (2001), A simple single-layer urban canopy model for atmospheric models: Comparison with multi-layer and slab

Page 36: Contribution of Impervious Surfaces to Urban ... - Princeton

models, Boundary-Layer Meteorology, 101(3), 329–358.

Lee, S.-H., and S.-U. Park (2008), A Vegetated Urban Canopy Model for Meteorological and Environmental Modelling, Boundary-Layer Meteorology, 126(1), 73–102.

Lemonsu, A., A. Leroux, S. Belair, S. Trudel, and J. Mailhot (2008), A general methodology of urban land cover type classification for atmospheric modelling, CRTI Scientific Report.

Lemonsu, A., V. Masson, and E. Berthier (2007), Improvement of the hydrological component of an urban soil–vegetation–atmosphere–transfer model, Hydrol. Process., 21(16), 2100–2111.

Leroyer, S., J. Mailhot, S. Bélair, A. Lemonsu, and I. B. Strachan (2010), Modeling the Surface Energy Budget during the Thawing Period of the 2006 Montreal Urban Snow Experiment, Journal of Applied Meteorology and Climatology, 49(1), 68–84.

Li, D., and E. Bou-Zeid (2011), Coherent structures and the dissimilarity of turbulent transport of momentum and scalars in the unstable Atmospheric surface layer, Boundary-Layer Meteorology, 140(2), 243–262.

Li, D., and E. Bou-Zeid (2013), Synergistic interactions between urban heat islands and heat waves: the impact in cities is larger than the sum of its parts, Journal of Applied Meteorology and Climatology, 52(9), 2051–2064.

Li, D., E. Bou-Zeid, and H. A. R. DeBruin (2012), Monin–Obukhov Similarity Functions for the Structure Parameters of Temperature and Humidity, Boundary-Layer Meteorology, 1–23.

Martilli, A., A. Clappier, and M. W. Rotach (2002), An Urban Surface Exchange Parameteterisation For Mesoscale Models, Boundary-Layer Meteorology, 104(2), 261–304.

Masson, V. (2000), A physically-based scheme for the urban energy budget in atmospheric models, Boundary-Layer Meteorology, 94(3), 357–397.

Masson, V., C. Grimmond, and T. R. Oke (2002), Evaluation of the Town Energy Balance (TEB) scheme with direct measurements from dry districts in two cities, Journal of Applied Meteorology, 41(10), 1011–1026.

Mills, G. M. (1993), Simulation of the energy budget of an urban canyon--I. Model structure and sensitivity test, Atmospheric Environment. Part B. Urban Atmosphere, 27(2), 157–170.

Mills, G. M., and A. J. Arnfield (1993), Simulation of the energy budget of an urban canyon—II. Comparison of model results with measurements, Atmospheric Environment. Part B. Urban Atmosphere, 27(2), 171–181.

Page 37: Contribution of Impervious Surfaces to Urban ... - Princeton

Monteith, J. L., G. Szeicz, and P. E. Waggoner (1965), The measurement and control of stomatal resistance in the field, Journal of Applied Ecology, 2(2), 345–355.

Noilhan, J., and S. Planton (1989), A Simple Parameterization of Land Surface Processes for Meteorological Models, Mon. Wea. Rev., 117(3), 536–549.

Nordbo, A., L. Järvi, S. Haapanala, and J. Moilanen (2013), Intra-city variation in urban morphology and turbulence structure in Helsinki, Finland, Boundary-Layer Meteorology, 146, 469–496.

Offerle, B., C. S. B. Grimmond, K. Fortuniak, and W. Pawlak (2006), Intraurban Differences of Surface Energy Fluxes in a Central European City, Journal of Applied Meteorology and Climatology, 45(1), 125–136.

Oke, T. R. (1982), The energetic basis of the urban heat island, Quarterly Journal of the Royal Meteorological Society, 108(455), 1–24.

Ramamurthy P., E. Bou-Zeid , J.A. Smith , Z. Wang, M.L. Baeck, N.Z. Saliendra, J. Hom, and C. Welty (2014) Influence of Sub-Facet Heterogeneity and Material Properties on the Urban Surface Energy Budget, Journal of Applied Meteorology and Climatology, in review.

Penman, H. (1948), Natural evaporation from open water, hare soil and grass, Proc R Soc Lond A Math Phys Sci, 193(1032), 120–145.

Richards, K. (2004), Observation and simulation of dew in rural and urban environments, Progress in Physical Geography, 28(1), 76–94.

Ross, P., and J. Parlange (1994), Comparing exact and numerical solutions of Richards' equation for one-dimensional infiltration and drainage, Soil science, 157(6), 341–344.

Rotach, M. W. et al. (2005), BUBBLE - a Major Effort in Urban Boundary Layer Meteorology, Theoretical and Applied Climatology, 81(3-4), 231–261.

Roth, M., and T. R. Oke (1995), Relative Efficiencies of Turbulent Transfer of Heat, Mass, and Momentum over a Patchy Urban Surface, 52, 1863–1874.

Rutter, A. J., K. A. Kershaw, P. C. Robins, and A. J. Morton (1972), A predictive model of rainfall interception in forests, 1. Derivation of the model from observations in a plantation of Corsican pine, Agricultural Meteorology, 9, 367–384.

Ryu, Y.-H., J.-J. Baik, and S.-H. Lee (2011), A New Single-Layer Urban Canopy Model for Use in Mesoscale Atmospheric Models, Journal of Applied Meteorology and Climatology, 50(9), 1773–1794.

Shepherd, J. M. (2005), A review of current investigations of urban-induced rainfall and recommendations for the future, Earth Interactions, 9(12), 1–27.

Page 38: Contribution of Impervious Surfaces to Urban ... - Princeton

Shuttleworth, W. J. (1978), A simplified one-dimensional theoretical description of the vegetation-atmosphere interaction, Boundary-Layer Meteorology, 14(1), 3–27.

Vesala, T., L. Järvi, S. Launiainen, A. Sogachev, Ü. Rannik, I. Mammarella, E. Siivola, P. Keronen, J. Rinne, and A. Riikonen (2008), Surface–atmosphere interactions over complex urban terrain in Helsinki, Finland, Tellus B, 60(2), 188–199.

Wang L., D. Li , Z. Gao, T. Sun, X. Guo, and E. Bou-Zeid (2014), Turbulent transport of momentum and scalars above an urban canopy, Boundary-Layer Meteorology, e-view, DOI: 10.1007/s10546-013-9877-z.

Wang, Z., E. Bou-Zeid, and J. A. Smith (2011a), A spatially-analytical scheme for surface temperatures and conductive heat fluxes in urban canopy models, Boundary-Layer Meteorology, 138(2), 171–193.

Wang, Z., E. Bou-Zeid, and J. A. Smith (2013), A coupled energy transport and hydrological model for urban canopies evaluated using a wireless sensor network, Quarterly Journal of the Royal Meteorological Society, 139(675), 1643–1657.

Wang, Z., E. Bou-Zeid, S. K. Au, and J. A. Smith (2011b), Analyzing the Sensitivity of WRF’s Single-Layer Urban Canopy Model to Parameter Uncertainty Using Advanced Monte Carlo Simulation, Journal of Applied Meteorology and Climatology, 50(9), 1795–1814.

Webb, E. K., G. I. Pearman, and R. Leuning (1980), Correction of flux measurements for density effects due to heat and water vapour transfer, Quarterly Journal of the Royal Meteorological Society, 106(447), 85–100.

Wilzack, J., S. Oncely, and S. Stage (2001), Sonic anemometer tilt correction algorithms, Boundary-Layer Meteorology, 99, 127–150.