Evaluating the calculated dry deposition velocities of reactive nitrogen oxides and ozone from two community models over a temperate deciduous forest

https://doi.org/10.1016/j.atmosenv.2011.02.063Get rights and content

Abstract

Hourly measurements of O3, NO, NO2, PAN, HNO3 and NOy concentrations, and eddy-covariance fluxes of O3 and NOy over a temperate deciduous forest from June to November, 2000 were used to evaluate the dry deposition velocities (Vd) estimated by the WRF-Chem dry deposition module (WDDM), which adopted Wesely (1989) scheme for surface resistance (Rc), and the Noah land surface model coupled with a photosynthesis-based Gas-exchange Evapotranspiration Model (Noah-GEM). Noah-GEM produced better Vd(O3) variations due to its more realistically simulated stomatal resistance (Rs) than WDDM. Vd(O3) is very sensitive to the minimum canopy stomatal resistance (Ri) which is specified for each seasonal category assigned in WDDM. Treating Sep-Oct as autumn in WDDM for this deciduous forest site caused a large underprediction of Vd(O3) due to the leafless assumption in ‘autumn’ seasonal category for which an infinite Ri was assigned. Reducing Ri to a value of 70 s m−1, the same as the default value for the summer season category, the modeled and measured Vd(O3) agreed reasonably well. HNO3 was found to dominate the NOy flux during the measurement period; thus the modeled Vd(NOy) was mainly controlled by the aerodynamic and quasi-laminar sublayer resistances (Ra and Rb), both being sensitive to the surface roughness length (z0). Using an appropriate value for z0 (10% of canopy height), WDDM and Noah-GEM agreed well with the observed daytime Vd(NOy). The differences in Vd(HNO3) between WDDM and Noah-GEM were small due to the small differences in the calculated Ra and Rb between the two models; however, the differences in Rc of NO2 and PAN between the two models reached a factor of 1.1–1.5, which in turn caused a factor of 1.1–1.3 differences for Vd. Combining the measured concentrations and modeled Vd, NOx, PAN and HNO3 accounted for 19%, 4%, and 70% of the measured NOy fluxes, respectively.

Highlights

► The first study on evaluating Vd(NOy) in the dry deposition models. ► Compared the performance of two models with different canopy treatments. ► Assessed the sensitivity of key parameters to simulate Vd(O3) and Vd(NOy). ► Improved the models by comparing with the field observations. ► Further developed the GEM model.

Introduction

Global atmospheric emissions of nitrogen oxide have increased dramatically during the past 150 years, and the supply of reactive nitrogen to ecosystems has doubled due to anthropogenic activities such as nitrogen fertilization, biomass burning, and fossil fuel combustion (Galloway et al., 2008). Dry deposition is responsible for a significant portion of the total (wet and dry) nitrogen deposition (e.g. 34%, Munger et al., 1998; 58%, Sparks et al., 2008). Up to 43% of NOx–N emissions over North America have been estimated to be removed from the atmosphere by dry deposition (Shannon and Sisterson, 1992). Reactive nitrogen oxides, called NOy, is a class of oxidized nitrogen compounds including NO, NO2, NO3, N2O5, HNO3, PAN (peroxyacetyl nitrate), other organic nitrates, and particle nitrate, which supply significant nutrient and acidic quantities to ecosystems. Augmented atmospheric deposition of NOy associated with increased emissions of NOx poses many environmental threats, including acidification of soil and surface water, eutrophication of lake, river and estuary, loss of biodiversity, damage to forests, and global climate change (Galloway et al., 2008). Increased anthropogenic emissions of NOx combined with hydrocarbons have produced high levels of surface O3 concentration. O3 can penetrate the tissues of leaves easily through stomatal uptake, causing stomatal occlusion and leaf damage. The direct uptake by vegetation through the stomata is also a major sink of O3 in the lower troposphere (Turnipseed et al., 2009).

Given the significant impacts of NOy and O3 deposition on atmospheric chemistry and ecosystem health, it is desirable to quantify the deposition amount and assess the effects. Measuring deposition fluxes for reactive nitrogen compounds and O3 with the eddy-covariance technique (e.g. Munger et al., 1996, Turnipseed et al., 2006) or the gradient method (e.g. Meyers et al., 1989, Sievering et al., 2001) have formed the basis for deposition models aimed at predicting dry depositions of reactive nitrogen compounds and O3.

Models have been developed (e.g. Wesely, 1989, Meyers et al., 1998, Zhang et al., 2002, Zhang et al., 2003, Niyogi et al., 2009, Wu et al., 2003) to estimate the dry deposition velocity (Vd) by commonly utilizing the resistance approach analogous to Ohm’s law in electrical circuits. Accurately parameterizing the complex surface-atmosphere exchange process remains challenging for Vd modeling due to large variability in surface conditions (e.g., vegetation types, and soil contents) at model sub-grid scales. It is difficult to fully describe the physiological processes concerning the vegetation stomatal responses to various environmental conditions, leaf age, injury, and so on. The rapid within-canopy chemical reactions are not often considered in simple single-layer models, neither for the role of horizontal flow to receptor surfaces over non-uniform surfaces and terrains (Wesely and Hicks, 2000). Therefore, large uncertainties still exist in modeling Vd. A recent study (Flechard et al., 2010) modeled the Vd of inorganic reactive nitrogen species (i.e. NH3, NO2, HNO3, and HONO and aerosol NH4+ and NO3) over 55 monitoring sites throughout Europe, using four existing dry deposition models. Their result revealed that differences between models can reach a factor 2–3 and are even greater than differences between monitoring sites. Hence, there is a continuous need to evaluate modeled Vd over different land-cover types and for different chemical compounds.

Observational deposition fluxes of SO2 and O3 are often used to evaluate models (Zhang et al., 2002, Wu et al., 2003). However, few studies have evaluated modeled Vd for nitrogen species primarily because accurate quantifications of dry deposition fluxes and speciation of the reactive nitrogen species are difficult and expensive to obtain (Horii et al., 2005). Munger et al. (1996) demonstrated that the dry deposition fluxes of NOy can be measured reliably using the eddy-covariance technique and year-round observations have been conducted at the Harvard Forest Environmental Measurement Site (HFEMS) since 1990. In a campaign attempting to estimate NOy concentration and deposition budget, concentrations of individual NOy species (i.e. NO, NO2, PAN and HNO3) have been measured at HFEMS. The reactive nitrogen dataset along with the O3 fluxes/concentrations available at HFEMS are used to evaluate two community dry deposition models here.

One model is the Weather Research and Forecasting-Chemistry model (WRF-Chem) dry deposition module (hereafter WDDM). WRF-Chem is a state-of-the-art, regional atmospheric chemistry model (Grell et al., 2005) and has been successfully applied for regional air quality studies (Wang et al., 2009). Due to lack of observational data, few studies have evaluated the ability of the WDDM for calculating nitrogen Vd, even though dry deposition is one of the most important sinks for pollutants. The other model is the Noah land surface model (LSM) (Chen and Dudhia, 2001) coupled with a photosynthesis-based Gas-exchange Evapotranspiration Model (Niyogi et al., 2009) (hereafter Noah-GEM). The Noah LSM has been used to provide surface heat fluxes as boundary conditions for WRF. It is of broad interest to develop capacities of computing Vd in Noah LSM (Charusombat et al., 2010). This evaluation effort is part of a broader effort to eventually integrate the balance of hydrosphere, biosphere, and atmosphere with environmental modeling such as atmospheric nitrogen input for the ecosystems in Noah. There are also plans to couple surface deposition and emission information more closely in Noah by linking with biogenic emission models such as MEGAN (Model of Emissions of Gases and Aerosols from Nature; Guenther et al., 2006). So one main purpose of this paper is to document current deficiencies in WDDM and raise the awareness of such problems. Also, because an investigation of nitrogen deposition calculation has not been done for these models, this study takes advantage of recently available nitrogen flux data to investigate nitrogen-deposition algorithms, which can serve well in the deposition models. The objectives are to: 1) assess the performances of WDDM and Noah-GEM in calculating Vd(NOy) and Vd(O3) over a temperate deciduous forest, 2) understand the sensitivity of modeled Vd(NOy) and Vd(O3) to the key variables/parameters, and 3) improve the models by comparing with the field observations.

We will first describe the measurements used in this study (Section 2) and the modeling framework and formulations of WDDM and Noah-GEM (Section 3). Next, the observation data and model results and discussions are presented in Section 4, which is followed by the conclusions in Section 5.

Section snippets

Site description

The HFEMS is located in a temperate 80–100 year-old mixed deciduous forest in central Massachusetts (42.54 N, 72.18 W; elevation, 340 m), which consists of red oak (Quercus rubra), red maple (Acer rubrum) with scattered hemlock (Tsuga canadensis), red pine (Pinus resinosa), and white pine (Pinus strobus). The canopy height near the observation tower is approximately 20 m with a peak leaf area index (LAI) of 3.4 m2 m−2 during summer. The nearest sources of significant pollution are a secondary road

Modeling framework

The resistance method determines Vd as the reciprocal of a total resistance (Rt) which consists of a series of resistances to perform gas transport from the atmosphere down to the surface.Vd(z)=Rt1=(Ra(z)+Rb+Rc)1Table A.1 describes each resistance component, and Table A.2 compares the formulations between WDDM and Noah-GEM.

Further developments of GEM

The GEM model (Niyogi et al., 2009) was further developed here (see Appendix B), but the parameters were kept the same and not specifically tuned for this study. Rs is the

The observations of O3 deposition and its environmental drivers

Fig. 2 shows the time series of hourly-averaged [O3] and F(O3) from June–October 2000. There was a distinct seasonal cycle of [O3] showing maxima in summer, associated with the high solar radiation and temperature. The peak values ranged from 40 to 80 ppbv, slightly lower than the observations in 1991–1994 (Munger et al., 1996). F(O3) followed the same seasonal trend with maxima during summer, closely coinciding with high concentrations and canopy growth (Munger et al., 1996). As shown in Fig. 3

Summary and conclusions

We evaluated the ability of two models (WDDM and Noah-GEM) to calculate Vd(O3) and Vd(NOy) against direct observations at HFEMS, and identified key variables/parameters and uncertainties in the two models. WDDM employs Wesely (1989) parameterization for Rc, which uses a simple Rs scheme based on the Ri parameter prescribed for each season and land-cover category. The uncertainty in prescribed Ri dominates the errors in estimating Vd for O3 and other gases that are controlled by the stomatal

Acknowledgements

This work was supported by the Natural Science Foundation of China (grant Nos. 40875076, U0833001), the National Program on Key Basic Research Project of China (973) (grant No. 2010CB428504) and the Fundamental Research Funds for the Central Universities. We also gratefully acknowledge the NCAR Advanced Study Program (ASP), BEACHON and Water System Programs. This research also benefited through the NOAA/JCSDA grant (NA06NES4400013), NASA-Terrestrial Hydrology Program and DOE ARM program. The

References (39)

  • J. Ball et al.

    A model predicting stomatal conductance and its contribution to the control of photosynthesis under different environmental conditions

  • U. Charusombat et al.

    Evaluating a new deposition velocity module in the Noah land surface model

    Boundary-Layer Meteorology

    (2010)
  • F. Chen et al.

    Impact of atmospheric surface layer parameterization in the new land-surface scheme of the NCEP mesoscale Eta numerical model

    Boundary-Layer Meteorology

    (1997)
  • F. Chen et al.

    Coupling an advanced land surface – hydrology model with the Penn State–NCAR MM5 modeling system. part I, model implementation and sensitivity

    Monthly Weather Review

    (2001)
  • F. Chen et al.

    On the coupling strength between the land surface and the atmosphere: from viewpoint of surface exchange coefficients

    Geophysical Research Letters

    (2009)
  • E.J. Cooter et al.

    Sensitivity of the National Oceanic and Atmospheric Administration multilayer model to instrument error and parameterization uncertainty

    Journal of Geophysical Research

    (2000)
  • C.R. Flechard et al.

    Dry deposition of reactive nitrogen to European ecosystems: a comparison of inferential models across the NitroEurope network

    Atmospheric Chemistry and Physics Discussions

    (2010)
  • J.N. Galloway et al.

    Transformation of the nitrogen cycle: recent trends, questions, and potential solutions

    Science

    (2008)
  • A. Guenther et al.

    Estimates of global terrestrial isoprene emissions using MEGAN, Model of Emissions of Gases and Aerosols from Nature

    Atmospheric Chemistry and Physics

    (2006)
  • Cited by (61)

    • Monitoring nitrogen deposition in global forests

      2023, Atmospheric Nitrogen Deposition to Global Forests: Spatial Variation, Impacts, and Management Implications
    • Assessment and intercomparison of ozone dry deposition schemes over two ecosystems based on Noah-MP in China

      2022, Atmospheric Environment
      Citation Excerpt :

      This may be because the morphological parameters for the vegetation used in the schemes were uniform both below and above the canopy, and this does not reflect the real forest ecosystem. According to Wu et al. (2011), the structure of the canopy plays a significant role in the simulated results, which may further account for this gap. As Fig. 2 shows, the schemes based on the Jarvis stomatal conductance mechanism showed a bimodal trend, reaching a peak at around 8:00–9:00 and 16:00–17:00.

    View all citing articles on Scopus
    View full text