COMMUNITY MULTISCALE AIR QUALITY (CMAQ) MODELING OF THE ATMOSPHERIC QUALITY AND POLLUTANT DEPOSITION OVER THE TERRACE-KITIMAT VALLEY OF NORTHWESTERN BRITISH COLUMBIA, CANADA by Chibuike Onwukwe B. Tech., Federal University of Technology Owerri, 2005 M. Sc., École Centrale de Nantes, 2015 DISSERTATION SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY IN NATURAL RESOURCES AND ENVIRONMENTAL STUDIES UNIVERSITY OF NORTHERN BRITISH COLUMBIA September 2020 © Chibuike Onwukwe, 2020 Abstract Tracking the effects of air pollution from industries is important for developing management strategies under changing emissions. However, computational tools for air pollution assessment often do not elucidate modeling uncertainty, making it difficult for environmental policy-makers to know how much confidence to put in model results, which also hampers aspects that may need improving. This study examined how the WRF-SMOKE-CMAQ modeling system with various planetary boundary-layer (PBL) schemes and atmospheric datasets mimics the local meteorology, air quality and acidic deposition at 1 km horizontal resolution over the industrializing Terrace-Kitimat Valley of northwestern British Columbia. Quantitative and qualitative correspondence of model outputs with observational data varied with station location, the nature of pollutant emissions, and quantity of chemical species. Valid model outputs were used to delineate present compliance with objectives on ambient fine particulate matter, and baseline exceedance of critical loads of sulfur and nitrogen deposition for the forest ecosystem. Spatial impacts of anticipated industrial emissions on the environment were also assessed. An additional 15 tonnes day-1 permissible SO2 emission from an aluminum smelter in Kitimat was projected to result in 50–88 % increase in aerial exceedance of the limit for protection of lichen, and 37–67 % increase in spatial exceedance of acidic deposition to soils. Cumulatively, 16–18 km2 of plant habitat, and 10–11 km2 of soil in an area contiguous with the smelter site will likely be damaged by its SO2 emission under the latest regulation. Should two Liquefied Natural Gas projects commence operations, cumulative NOx concentrations are expected to remain below harmful levels, while pre-existing areal exceedance of nitrogen deposition will barely increase (0–1 km2). An additional 4 km2 area will be exposed to SO2 concentrations ii that are directly harmful to vegetation, while 13–14 km2 total area with an average of 29.7–35.0 kg ha-1 yr-1 excess sulfur deposition was estimated. These projections assumed all future emissions of NOx, SO2 and other air pollutants will be from elevated point sources. iii Preface This dissertation contains the original research and analyses conducted by the author, Chibuike Onwukwe, under the guidance of Peter Jackson. It contains many figures, some of which are maps. All maps were created by the author using various software such as NCL (http://www.ncl.ucar.edu/), google maps javascript (https://developers.google.com/ maps/documentation/javascript/tutorial?), QGIS (http://www.qgis.org/en/site/), and map viewer of the Integrated Land and Resource Registry, British Columbia Ministry of Forests, Lands, Natural Resource Operations and Rural Development. Chapter 2 of this dissertation is published in Volume 59, Issue 8 of J. Appl. Meteorol. Climatol., Chapter 3 is published in Volume 233 of Atmos. Environ., Chapter 4 is in Press at the J. Air Waste Manage. Assoc., while Chapter 5 is published in Atmos.Pol. Res. iv Table of Contents Page Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii Acronyms, Notations & Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xx 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1 Geo-ecological imperatives for atmospheric pollution research: an overview 1 1.2 Motivation for present study . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.3 Study area background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.4 Research questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 1.5 Research Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2 Meteorological downscaling with WRF model version 4.0 and comparative evaluation of planetary boundary layer schemes over a complex coastal airshed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 2.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 2.2.1 Description of the WRF model and PBL schemes . . . . . . . . . . 16 2.2.2 Domain configuration, model initialization and settings . . . . . . 19 2.2.3 Observational data . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.2.4 Statistical evaluation indices . . . . . . . . . . . . . . . . . . . . . . 22 2.3 Results and analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 v 2.3.1 Synoptic and mesoscale representations . . . . . . . . . . . . . . . 23 2.3.2 Performance evaluations for surface variables . . . . . . . . . . . . 27 2.3.3 Spatio-differential examinations of surface meteorological fields . 34 2.4 Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 3 Evaluation of CMAQ modeling sensitivity to planetary boundary layer parameterizations for gaseous and particulate pollutants over a fjord valley . 50 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 3.2 Methods and data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 3.2.1 PBL schemes and experiments set-up . . . . . . . . . . . . . . . . . 53 3.2.2 Air quality monitoring data . . . . . . . . . . . . . . . . . . . . . . 56 3.2.3 Statistical evaluation indices . . . . . . . . . . . . . . . . . . . . . . 57 3.3 Results and analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 3.3.1 Ambient air quality . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 3.3.2 Model performance evaluations . . . . . . . . . . . . . . . . . . . . 63 3.3.3 Annual evaluations and spatial concentration differences . . . . . 73 3.4 Discussion and conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 4 Gridded bias correction of modeled PM 2.5 for exposure assessment and estimation of background concentrations over a coastal valley in northwestern British Columbia, Canada . . . . . . . . . . . . . . . . . . . . . . . . . 83 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 4.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 4.2.1 Model simulations and observations data . . . . . . . . . . . . . . 86 4.2.2 Bias correction formulations . . . . . . . . . . . . . . . . . . . . . . 88 4.2.3 Evaluation measures . . . . . . . . . . . . . . . . . . . . . . . . . . 88 4.3 Results and analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 4.3.1 Accuracy of bias corrected outputs . . . . . . . . . . . . . . . . . . 90 4.3.2 Evaluations for fitness with compliance metrics . . . . . . . . . . . 92 4.3.3 Estimation of background concentrations . . . . . . . . . . . . . . 96 4.4 Discussion and conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 5 Acid wet-deposition modeling sensitivity to WRF-CMAQ planetary boundary layer schemes and exceedance of critical loads over a coastal mountain valley area of northwestern British Columbia, Canada . . . . . . . . . . . . . 105 vi 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 5.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 5.2.1 Deposition simulations . . . . . . . . . . . . . . . . . . . . . . . . . 110 5.2.2 Measurement data and performance measures . . . . . . . . . . . 111 5.3 Results and analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 5.3.1 Performance evaluation for wet deposition of acidifying species . 114 5.3.2 Comparison of total nitrogen and sulfur deposition among PBL schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 5.3.3 2017 critical-load exceedances for forest ecosystems . . . . . . . . 124 5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 5.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 6 Intercomparison of atmospheric datasets and PBL schemes for precipitation downscaling over a coastal mountain valley of northern British Columbia, Canada . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 6.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 6.2.1 Description of datasets . . . . . . . . . . . . . . . . . . . . . . . . . 138 6.2.2 Modeling configurations and experiments . . . . . . . . . . . . . . 139 6.2.3 Observational data and evaluation protocol . . . . . . . . . . . . . 141 6.3 Results and analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143 6.3.1 Quantitative biases for datasets and PBL schemes . . . . . . . . . . 143 6.3.2 Spatio-temporal verifications . . . . . . . . . . . . . . . . . . . . . . 147 6.3.3 Predictive evaluations versus summary distributions . . . . . . . 151 6.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 6.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 7 Modeling terrestrial ecosystems exposure to incremental smelter sulfur dioxide emissions and deposition in a complex coastal valley airshed . . . . 162 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 7.2 Modeling framework and procedure . . . . . . . . . . . . . . . . . . . . . 166 7.3 Results and analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168 7.3.1 Baseline SO 2 levels and changes . . . . . . . . . . . . . . . . . . . . 168 7.3.2 Critical level exceedances and mapping . . . . . . . . . . . . . . . 170 7.3.3 Critical load of acidity exceedances and mapping . . . . . . . . . . 172 vii 7.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 7.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 8 Quantifying incremental and cumulative terrestrial ecosystems impacts of NOx and SO2 emissions from LNG operations in the Terrace-Kitimat valley of northwestern British Columbia . . . . . . . . . . . . . . . . . . . . . . . 181 8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182 8.2 LNG emissions and numerical modeling set-up . . . . . . . . . . . . . . . 185 8.3 Results and analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188 8.3.1 Relative changes in ambient NO x and SO2 and exceedance of critical levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188 8.3.2 Relative changes in nitrogen and sulfur deposition and exceedance of critical loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192 8.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 8.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201 9 Summary and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203 9.0.1 Recapitulation of research findings . . . . . . . . . . . . . . . . . . 203 9.0.2 Significance of study findings . . . . . . . . . . . . . . . . . . . . . 208 9.0.3 Directions for future research . . . . . . . . . . . . . . . . . . . . . 210 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211 Appendix A Selection of meteorological year for model simulations . . . . . . . . . . . . 243 B WRF physics settings and model heights . . . . . . . . . . . . . . . . . . . . . 246 C Formulae for statistical measures . . . . . . . . . . . . . . . . . . . . . . . . . 248 D Verification of atmospheric sounding . . . . . . . . . . . . . . . . . . . . . . . 249 E Summer and winter days comparisons of air temperature and wind speed to seasonal averages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250 F Performance statistics for simulation of surface meteorological variables . . 252 G Centreline profile profiles of key WRF output variables . . . . . . . . . . . . 257 viii H Formulae for indices of precipitation prediction . . . . . . . . . . . . . . . . 259 I The implication of sulfur exceedance on vegetated land . . . . . . . . . . . . 260 ix List of Tables 2.1 Properties of WRF domains with ERA5 . . . . . . . . . . . . . . . . . . . . . 21 2.2 Stations for verification of WRF-ERA5 downscaling . . . . . . . . . . . . . . 22 2.3 Performance benchmarks for surface meteorological variables . . . . . . . . 23 3.1 WRF-SMOKE-CMAQ domains attributes . . . . . . . . . . . . . . . . . . . . 56 3.2 Air quality stations having valid data in 2017 . . . . . . . . . . . . . . . . . . 57 3.3 Benchmarks for air quality modeling . . . . . . . . . . . . . . . . . . . . . . . 58 3.4 Averages of observed pollutant concentrations in peak seasons . . . . . . . . 59 3.5 Modeling performance: SO 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 3.6 Modeling performance: PM2.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 3.7 Modeling performance: NO2 3.8 Modeled annual peak concentrations of air pollutants . . . . . . . . . . . . . 76 4.1 PM2.5 monitoring data completeness . . . . . . . . . . . . . . . . . . . . . . . 87 4.2 Statistical evaluation of bias-corrections compared to original model output 4.3 Accuracy of bias corrected model outputs for compliance indicators . . . . . 93 4.4 Tiers of air quality management based on ambient PM2.5 . . . . . . . . . . . 94 . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 x 91 5.1 Model performance statistics for wet deposition of ammonium, nitrate and sulfate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 5.2 Percentage contribution of wet deposition in 2017 to S and N annual deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 5.3 Exceedances of critical loads of acidity (CLA) and nutrient nitrogen . . . . . 127 6.1 Nesting with NAM_ANL versus ERA5/ NARR . . . . . . . . . . . . . . . . 140 6.2 Observational data locations for evaluation of precipitation simulations . . . 142 6.3 Fitness measures for precipitation simulations . . . . . . . . . . . . . . . . . 143 6.4 Percent bias for 2017 precipitation simulation at station locations . . . . . . . 145 6.5 Contributions of snow water equivalent to total precipitation in observation and simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 6.6 Predictive scores of daily precipitation for all events ( ≥ 0.2 mm) and events ≥ mean amounts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153 7.1 Meteorological and emissions domain sizes and attributes for SMOKE-CMAQ modeling with ERA5 and NAM_ANL data . . . . . . . . . . . . . . . . . . . 168 7.2 Estimates of aerial SO2 exceedance of critical levels of vegetation and lichen exposures due to smelter emissions changes . . . . . . . . . . . . . . . . . . . 173 7.3 Estimates of sulfur deposition exceedance from aluminum smelter’s SO2 emissions in Kitimat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 8.1 Estimated exceedances of critical level (3.6 ppb) of SO2 lichen exposure with, and without LNG emissions 8.2 . . . . . . . . . . . . . . . . . . . . . . . . 192 Estimated exceedances of critical load of soil nutrient nitrogen with, and without LNG emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194 xi 8.3 Estimated exceedances of critical load of sulfur with, and without LNG emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197 A.1 Ranking of differences for meteorological year selection . . . . . . . . . . . . 244 B.1 Settings for WRF physics/dynamics . . . . . . . . . . . . . . . . . . . . . . . 246 B.2 WRF model layer heights . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247 F.1 Performance statistics of WRF simulation of air temperature . . . . . . . . . 253 F.2 Performance statistics of WRF simulation of specific humidity . . . . . . . . 254 F.3 Performance statistics of WRF simulation of wind direction . . . . . . . . . . 255 F.4 Performance statistics of WRF simulation of wind speed . . . . . . . . . . . 256 H.1 Contingency table for precipitation events forecast . . . . . . . . . . . . . . . 259 xii List of Figures 1.1 Description of study area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 1.2 Surficial geology and landcover of the TKV area . . . . . . . . . . . . . . . . 7 1.3 Land use and development in the TKV . . . . . . . . . . . . . . . . . . . . . . 8 2.1 WRF nesting with ERA5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.2 Simulated summertime meteorological fields . . . . . . . . . . . . . . . . . . 26 2.3 Simulated wintertime meteorological fields . . . . . . . . . . . . . . . . . . . 28 2.4 Diurnal air temperature bias with ERA5 . . . . . . . . . . . . . . . . . . . . . 29 2.5 Diurnal wind speed bias with ERA5 . . . . . . . . . . . . . . . . . . . . . . . 33 2.6 Spatial differences amongst PBL schemes for meteorological variables: nighttime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 2.7 Spatial differences amongst PBL schemes for meteorological variables: daytime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 2.8 Vertical water vapor profiles for simulations with various PBL schemes . . . 44 2.9 Vertical profiles of potential temperature for simulations with various PBL schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 2.10 Effect of altenative surface-layer schemes on the MYNN3 PBL scheme . . . 49 3.1 WRF-SMOKE-CMAQ domains for ERA5 simulations . . . . . . . . . . . . . 55 xiii 3.2 Observed meteorology relevant to pollutant dispersion . . . . . . . . . . . . 61 3.3 Wind dependency of air pollutants . . . . . . . . . . . . . . . . . . . . . . . . 62 3.4 SO2 modeling evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 3.5 PM2.5 modeling evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 3.6 NO2 modeling evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 3.7 Pollutant modeling evaluation of concentration quantiles . . . . . . . . . . . 75 3.8 Spatial distribution of modeled annual concentrations of air pollutants . . . 77 4.1 PM2.5 bias correction formulae and quantile plots of modeled and observed PM2.5 concentrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 4.2 PM2.5 evaluation for categorical correspondence . . . . . . . . . . . . . . . . 95 4.3 Spatial plot of bias-corrected annual PM2.5 concentrations . . . . . . . . . . . 97 4.4 Background PM2.5 estimation from bias-corrected centreline concentrations 4.5 Classification of exposure to outdoor PM2.5 in 2017 . . . . . . . . . . . . . . . 102 5.1 Nesting set-up for WRF-CMAQ deposition modeling . . . . . . . . . . . . . 112 5.2 Weekly time series of observed wet deposition . . . . . . . . . . . . . . . . . 115 5.3 Spatial distribution of annual total nitrogen and sulfur deposition . . . . . . 120 5.4 Pairwise differences in annual total nitogen and sulfur deposition between 98 PBL schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 5.5 Domain-wide contribution of wet deposition as a ratio of total deposition . 123 5.6 Exceedances of critical loads of acidity and nutrient nitrogen . . . . . . . . . 126 6.1 WRF nesting with NAM_ANL versus ERA5/ NARR 6.2 Distribution of monthly precipitation in 2017 . . . . . . . . . . . . . . . . . . 145 xiv . . . . . . . . . . . . . 141 6.3 Spatial plots of dataset-normalized annual precipitation estimates . . . . . . 146 6.4 Spatial plots of PBL schemes-normalized annual precipitation estimates . . 147 6.5 Monthly series of observations (OBS) at gauge stations, and simulations with the MYJ PBL scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 6.6 Pairwise stations-normalized 2017 precipitation . . . . . . . . . . . . . . . . 150 6.7 Pairwise station-normalized 2017 precipitation . . . . . . . . . . . . . . . . . 152 6.8 Categorical evaluation of simulated precipitation amounts by various datasets154 6.9 Cumulative mean distributions of simulated and observed daily precipitation155 7.1 SO2 concentrations distribution and relative increase . . . . . . . . . . . . . . 170 7.2 SO2 spatial exceedance for lichen in Kitimat . . . . . . . . . . . . . . . . . . . 172 7.3 Modeled sulfur deposition exceedance from aluminum smelter’s SO2 emissions in Kitimat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174 7.4 Annual mean concentrations from maximum SO2 emissions of 42 tonnes day−1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 8.1 Locations of two LNG projects in the Kitimat area . . . . . . . . . . . . . . . 187 8.2 Annual emission estimates of major air pollutants from proposed LNG projects in Kitimat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188 8.3 NO x and SO2 concentration changes relative to levels without contributions from the LNG industry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 8.4 Exceedance of critical level (3.6 ppb) of lichen exposure to SO2 with, and without contributions from LNG industry emissions . . . . . . . . . . . . . . 191 8.5 Nitrogen and sulfur deposition changes relative to the loads without contributions from LNG industry emissions . . . . . . . . . . . . . . . . . . . . . 193 xv 8.6 Modeled spatial exceedance of critical load of acidity with, and without contributions from LNG industry . . . . . . . . . . . . . . . . . . . . . . . . . 196 8.7 Along-valley NO x and SO2 concentrations from cumulative emissions . . . 198 A.1 Observed wind speed, precipitation and air temperature in 2017 versus the averages for 2006–2018 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245 D.1 Verification of regional atmospheric sounding . . . . . . . . . . . . . . . . . . 249 E.1 Winterday and summerday comparisons to observed seasonal averages of air temperature and wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251 G.1 Valley centreline profiles of WRF diagnosed meteorological variables . . . . 258 xvi Acronyms, Notations & Units Acronyms ACM2 Asymetric Convective Model version 2 BC British Columbia BCMOECS BC Ministry of Environment and Climate Change Strategy CCME Canadian Council of Ministers of the Environment CMAQ Community Multiscale Air Quality ECCC Environment and Climate Change Canada ECMWF European Centre for Medium-range Weather Forecasts ERA5 ECMWF fifth major global reanalyses FAC2 Modeled values within a factor of 2 of observations LNG Liquefied natural gas MM5 Fifth-generation Penn State/NCAR mesoscale model MYJ Mellor–Yamada–Janjić MYNN3 Mellor–Yamada–Nakanishi–Niino level 3 NAM-ANL North American Mesoscale Analyses NARR North American Regional Reanalysis xvii NOx Nitrogen oxides NO2 Nitrogen dioxide PBL Planetary boundary layer PM2.5 Fine particulate matter PST Pacific Standard Time SH Shin–Hong SMOKE Sparse Matrix Operator Kernel Emissions SO2 Sulfur dioxide SST Sea surface temperature TKE Turbulent kinetic energy TKV Terrace–Kitimat Valley US United States USEPA US Environmental Protection Agency UTC Universal Time Coordinated UW University of Washington VOC Volatile organic compounds WHO World Health Organization xviii WRF Weather Research and Forecasting YSU Yonsei University Notations CLA Critical load of acidity CLNnut Critical load of nitrogen deposition NH4 + Ammonium NO3 − Nitrate SO4 2− Sulfate Units g kg-1 grams per kilogram kg ha −1 yr−1 kilograms per hectare per year m s-1 meters per second µg m-3 micrograms per cubic meter ppb parts per billion xix Acknowledgements I use this opportunity to express my deep appreciation to Peter Jackson. Exceptionally beneficient, Peter’s handling of students’ concerns is admirable. As supervisor, his guidance, candor, affableness, and absolute support saw to the completion of this study, and for this, I am eternally indebted. My immense gratitude extends to other members of the Supervisory Committee: Bruce Ainslie through whom I acquired the technical skills to approach this subject, Stephen Déry for providing critical feedbacks, and Mark Groulx for his insightful perspectives all the way. Collectively, I thank you for your time. I sincerely acknowledge financial awards for research and conference travels received from the UNBC Office of Graduate Studies. Reminiscing now, those travels served opportunities for meeting and exchanging ideas with atmospheric scientists from near and far, much of which proved valuable to my work. They also offered precious, albeit fleeting moments of relief, sustaining my health for which I am grateful. Much thanks goes to contacts in Kitimat / Terrace through whom local, experiential information on air quality was tapped, especially Steven and Lis Stannus and Pam Vollrath of the Kitimat–Terrace Air Quality Coalition, and Shawn Zettler of Rio Tinto Alcan. The assistance from you all is cherished. Lastly, I would like to thank colleagues with whom I shared office at some point: Rachel Hay and Brayden Nilson, neighbors: Aseem Sharma, Hadleigh Thompson and Nazrul Islam; and Prince Geoerge resident, Janelle Wright. Your company made studies bearable and I wish each one of you fulfilling careers. xx Chapter One Introduction 1.1 Geo-ecological imperatives for atmospheric pollution research: an overview Shared access to the atmosphere for disposal of gaseous wastes has resulted in the potential for unintended adverse impacts on ecosystems and societies. Apart from public health concerns (Kurt et al. 2016), air pollutants have been linked to harm on vegetation and soils. Sulfur dioxide (SO2 ) can induce bleaching and injury on foliage (Legge and Krupa 2002). High intake of nitrogen oxides (NOx ) by plants can reduce photosynthesis and yield (Hu et al. 2015). Exposure of trees to mixtures of several air pollutants can impair their resistance to stresses such as frost and snow storms (Vorobeichik et al. 2014). Further, excessive deposition of sulfur- and nitrogen-containing air pollutants to soils can lead to acidification, and nutrient enrichment (Augustaitis 2011), subsequently causing changes in composition of natural vegetation (Bobbink et al. 2010, Gilliam 2019). The severity of ecological effects of air pollution varies with plant species and habitats. Long 1 periods of cold temperatures and wet climates with frequent fog, low clouds and low light conditions can enhance the toxicity of gaseous air pollutants for flora (Augustaitis 2011, WHO 2000). Natural recovery of pollution-impacted ecosystems can be slow, and original plant communities may remain absent long after remedial measures are implemented (Vorobeichik et al. 2014). To prevent deleterious consequences on the environment, various jurisdictions have focused on inventorying anthropogenic air pollutants, and targeting large contributors for reduction. Commitments to large-scale reductions in emissions that contribute to crossborder acidifying pollution such as the 1991 Canada–U.S. Air Quality Agreement (ECCC 2018a), have complemented national and provincial actions to mitigate ecosystem degradation. At the core of measures for long-term management of acid rain in Canada (CCME 2014b), is monitoring of precursor atmospheric pollutants, and prescribing and enforcing ambient air quality standards. Airshed planning and management strategies are often based on data from pollution monitoring stations. 1.2 Motivation for present study Air pollutant levels could be very variable in space, thereby limiting the usefulness of fixed-site monitors. Even in the present era of low-cost sensors, it is inconceivable to have fixed monitors at every location. Reliable and accurate ground-based monitoring can be challenging and costly to set-up and operate for long periods especially in mountainous regions. Atmospheric-chemistry models such as CMAQ, that are capable of generating 2 gridded concentrations of a pollutant can, if shown to be sufficiently realistic, alleviate the challenge of field monitoring. These computational tools model the fate of pollutant releases from all known anthropogenic and biogenic sources, including their advection, chemical and physical transformation, and loss. They are suitable for multi-pollutant scenarios but are especially useful for accounting for pollution by species that are not directly released (secondary pollution). By simulating the sophisticated oxidant chemistry in the deposition of acidifying species, atmospheric chemistry models are increasingly being used to qualify habitat exposures to contaminants, thereby providing an integrated approach for projecting the effect of air emissions at various spatio-temporal scales. In addition, they are cost-saving since they can be used to assess alternative emissions scenarios which otherwise would be difficult to perform in the real world. Retrospective analysis can also be accomplished with them to provide understanding of atmospheric phenomena driving high pollutant concentrations. For these reasons, atmospheric chemistry models are widely accepted science tools for airshed planning and management. Canadian applications of atmospheric-chemistry models for acid deposition assessment (e.g Moran et al. 2008) ,however, have mainly been over parts of the eastern provinces where extensive harm to natural habitats attributable to both the lack of natural alkalinity of bedrock material, and to historically significant SO2 and NOx emissions from within and nearby, occur. The few modeling studies in Western Canada have either been over large areas of mostly flat terrain (e.g. Makar et al. 2018) or for wide, deindustrialized basins (e.g. Nasr et al. 2010). Coastal parts of British Columbia have low acid-buffering soils (ECCC 1991). Consequently, SO2 and NO x increase in places with 3 complicated topography could leave these terrestrial systems vulnerable to impacts of acid deposition. Modeling the effect of acid precursor release from industrial activity in fjord topography—areas with long, narrow coastal inlets bounded by steep mountains— presents unique needs that include discretizing the circulation in deep, narrow valleys. Accordingly, the Terrace-Kitimat Valley (TKV), an industrializing corridor of northwestern BC, serves an ideal testbed for atmospheric research. 1.3 Study area background Aligned north-south within the Pacific Coast mountains and flanked by ridges up to 1800 m high, the TKV lies at the head of a 90-km fjord (Fig. 1.1). Roughly 70 km in length, the valley is broadest around the small city of Terrace, but as narrow as 8 km at some sections further south, near the port town of Kitimat. The valley’s weather is dictated by air masses originating over both the Pacific Ocean and continental North America. In summer (June–August), when the Pacific High lingers near its northernmost position (Klock and Mullock 2000), clear, fine weather prevails. Surface wind is mostly southerly from the Douglas Channel, and land-sea breezes and slope-valley winds are experienced. In winter (December-–February), land-falling mid-latitude cyclones bring heavy snowfall and the intermittent clash of their onshore winds and northerly outflowing continental arctic air result in stormy weather (Lange 2003). The mean annual temperature in Kitimat (Terrace) is 6.9 °C (7.4 °C), varying from a minimum of -4 °C average in January to a maximum of + 22 °C average in July; annual precipitation amount is 2210 mm in Kitimat, but 1170 mm in Terrace (ECCC 2019), in the partial rain shadow of the Coast Mountains. Within 4 and over surrounding mountains are wilderness areas, comprising coniferous forests supported on podzolic soils (Fig. 1.2) (Soil Classification Working Group 1998). Vegetation on the valley floor, portions of which are drained by the Skeena River and Kitimat River, is dominated by western hemlock, western red cedar, Sitka spruce, Douglas-fir, amabilis fir; and mountain hemlock occurring at sub-alpine height (Pojar et al. 1991). Understorey abundance of bryophytes (Meidinger 2018) is prevalent. About 28,000 people live in the TKV, of which ∼ 12,000 are in and around the small city of Terrace, 8,300 in the port town of Kitimat, 600 in the First Nations community of Kitamaat Village, with the remainder scattered elsewhere (Statistics Canada 2018). Anthropogenic emissions in the valley are from these residential clusters, and light road and rail traffic, as well as industrial sources (Fig. 1.3). An existing aluminum smelter in Kitimat emits large amounts of SO2 . Re-emerging in 2016 after expansion/modernization works that boosted production (Rio Tinto 2016), daily permissible SO2 however, increased from 27 tonnes to 42 tonnes. Exports processing facilities such as liquefied natural gas plants (District of Kitimat 2020), have also been proposed in the area, some of which have been endorsed by the host Haisla and Tsimshian nations. Proposed developments however, have raised concerns about the air quality and deterioration of the natural environment (Environmental Appeal Board 2015). Projects when operational, will emit atmospheric pollutants (SO2 , NO x , particulate matter, etc.) in Kitimat. Pollutant modeling is needed, not only to determine present impacts but also, to project incremental and cumulative effects of anticipated emissions. 5 Figure 1.1 The Terrace–Kitimat Valley and surroundings, indicating the monitoring network for surface meteorology, air quality and atmospheric deposition. On the right is an enlargement of the Kitimat area, where an existing aluminum smelter (at the location of the industry symbol), about 2 km south of the Haul Road station, emits large amounts of SO2 . The inset on the left identifies the TKV in the Coast mountain ranges of northwestern British Columbia 6 Figure 1.2 Surficial geology of the Terrace–Kitimat Valley and surroundings (left), and land cover distribution (right) 1.4 Research questions This study investigates the impacts of air pollutant emissions, transport and deposition in the Terrace-Kitimat Valley and the surrounding area using the WRF-SMOKE-CMAQ modeling system (more details in subsequent chapters) at fine spatial resolution. Specific research objectives are highlighted in the following questions: 7 8 Figure 1.3 Land use and development in the TKV 1. What explains the behaviors of alternative PBL schemes in the WRF model for simulating the TKV’s surface meteorology? In numerical weather models, the prediction of phenomena that are too small to be resolved at the grid scale, such as turbulence are via various simplified formulae called parameterization schemes. Those for the planetary boundary layer (PBL) could impact the simulation of meteorological variables that are input to chemical transport models. PBL schemes are often developed under ideal conditions: flat terrain, and dry, warm environments (e.g. Hong et al. 2006, Nakanishi and Niino 2009). The applicability of PBL schemes in the very complex TKV area is unknown, therefore, should be evaluated. 2. How does the choice of PBL scheme affect the simulation of PM2.5 , SO2 and NO2 concentrations and what reasons may account for differences in model outputs? An assessment of the modeling skill of various PBL schemes for pollutant measurements in the TKV is conducted. This is necessary since modules in coupled numerical modeling systems are periodically updated and it is in the interest of the science community that revisions are rigorously evaluated. 3. Could quantile-based bias correction of CMAQ output improve usefulness for assessing regulatory compliance to air quality objectives? In cases where simulations have large biases, identifying whether post-processing of original outputs render them valid for assessing violations of standards on local air quality is important. 4. How well do simulations with various PBL parameterizations capture wet deposition 9 of acidifying ions and what are the baseline exceedances of sulfur and nitrogen deposition in the TKV area? Simulations of wet deposition which is a component of total deposition of acidifying species, is evaluated by comparing outputs from CMAQ to those of observational data at two measurement locations in the TKV. The credibility of outputs for various model runs are examined prior to estimating deposition impacts of existing emission sources in the valley. 5. What choices of atmospheric forcing data and PBL scheme in the WRF model can reproduce precipitation for the purpose of projecting atmospheric deposition over the TKV and what is the uncertainty in the precipitation field with the best fit simulations? Determining the validity of hydrological outputs of meteorological datasets, as substitutes of actual precipitation measurements, is needed for realistic projection of deposition impacts of future industrial emissions. 6. What incremental and aggregate impacts could arise should the existing smelter at Kitimat emit 42 tonnes day−1 that is the highest permissible rate? Although permitted up to 42 tonnes day−1 SO2 release, smelter emissions as at 2017 was still around the previous limit of 27 tonnes day −1 (ECCC 2020). Hence determining what pollutant concentrations, and incremental effects that may arise should the smelter emit at the maximum permissible amount, is crucial to understanding the sensitivity of the valley’s atmosphere to pollutant emissions. 7. For proposed developments in the TKV, 10 (a) To what additional extent could vegetation be exposed to harmful NOx and SO2 concentrations from LNG projects? (b) How much direct vegetation exposure to harmful NOx and SO2 concentrations could result from aggregate industrial emissions? (c) By how much could soil nitrogen enrichment and acidification change as a result of the LNG industry? (d) What aerial exceedances of critical loads of nitrogen and sulfur deposition will arise from aggregate industrial emissions? 1.5 Research Outline The rest of the dissertation consist of chapters, each of which address the research questions under topical headlines and in the same order as listed above, are in the form of manuscripts for peer-reviewed journals. Chapter 2 examines the influence of turbulence closures built-in to the WRF model in generating representative meteorological fields for the TKV area. It evaluates their emulation of surface meteorological parameters that affect air quality (e.g. air temperature, specific humidity and wind) and analyzes differences in outputs. Chapter 3 evaluates the capability of the WRF-SMOKE-CMAQ modeling system that is configured with various PBL schemes, to mimic observed air quality measurements in the valley (fine particulate matter, sulfur dioxide and nitrogen dioxide). It ranks the fitness of model runs both quantitatively and qualitatively, also discusses aspects for improve11 ment. Chapter 4 addresses the correction of bias in CMAQ output, and assesses whether improvement is appropriate for benchmarking exposure to ambient pollutant concentrations. Chapter 5 investigates wet acidic deposition sensitivity to PBL scheme choices. It also quantifies the uncertainty in exceedances of critical loads of nitrogen and sulfur deposition caused by base emission in TKV, prior to precursor emissions changes from planned industrial facilities. Chapter 6 assesses the credibility of different choices of atmospheric forcing data and PBL schemes for precipitation modeling, such that could affect the accuracy of acid deposition projections for the TKV area. Chapter 7 projects the implication of smelter SO2 emissions increase for vegetation exposure to ambient concentrations, and the risk of indirect impacts via soil acidification. Chapter 8 estimates the aerial impact of future export processing industries in the valley, specifically from planned liquefied natural gas facilities and allied shipping activities. Effects on atmospheric quality are modeled and evaluated from a cumulative concentrations/deposition assessment perspective. Chapter 9 provides an overall summary of the study’s findings and concludes the report. 12 Chapter Two Meteorological downscaling with WRF model version 4.0 and comparative evaluation of planetary boundary layer schemes over a complex coastal airshed (Also published at https://doi.org/10.1175/JAMC-D-19-0212.1 in J. Appl. Meteorol. Climatol.) Abstract Evaluation of downscaled meteorological information is crucial to identifying model behaviors that may propagate to end-applications such as the simulation of local air quality. This study conducted and assessed year-long simulations of hourly meteorology over the Terrace–Kitimat Valley of northwestern British Columbia at 1-km horizontal gridding for six PBL schemes in the WRF model, version 4.0. In terms of key surface meteorologi13 cal variables that affect air quality, simulations over land demonstrated better skill for specific humidity and wind direction, than for air temperature and wind speed. Spatial differences in modeled atmospheric properties and vertical profiles, especially for moisture content, were used to diagnose the relative capacity of each PBL scheme to represent pollution dispersion and dilution. Stable conditions at night increased suppression of boundary-layer mixing by the nonlocal YSU scheme compared with suppression by the local eddy-diffusion component of the ACM2 scheme, resulting in decreased wind speed and ambient temperature, but moister air with the YSU scheme. The weakening of mixing by the MYNN3 scheme with inland distance suggested that higher-order, nonlocal transport is sensitive to increasing topographic steepness towards the northern part of the valley. Disparities in mixing strengths amongst PBL schemes were greater in the summer when conditions were generally less stable with moist, warm air blowing inland than in winter when the valley channels cold, stable air from the interior. Increased convection in daytime led to greater entrainment of air from aloft and a thicker PBL with the YSU scheme than the ACM2 scheme in summer while increasing counter-gradient transport in the MYNN3 scheme that reduces dilution. 2.1 Introduction Correct representation of the wind, temperature, moisture and mixing heights are needed to predict the fate of air pollutants. It is generally recognized that PBL processes impact the prediction of these variables and many investigations of the sensitivity of coupled numerical models to PBL schemes has been undertaken in the last decade. Cheng et al. 14 (2012) simulated ozone episodes over Taiwan and found that discrepancies between results were due to vertical mixing strength and entrainment properties of the tested PBL schemes. Banks and Baldasano (2016) evaluated how outputs from some PBL schemes in the Weather Research and Forecasting (WRF) model differ from one deployed for operational air quality forecasts over Catalonia, Spain. In India, Gunwani and Mohan (2017) researched the sensitivity of PBL schemes to different climatic zones based on model simulation of some meteorological parameters. Others (e.g. Reboredo et al. 2015, Xie et al. 2012) have quantified meteorological prediction differences to ascertain which PBL scheme is most appropriate to specific geographical domains. However, much of the related literature (e.g. Mohan and Gupta 2018, Misenis and Zhang 2010) are for short episodes and do not portray cross-temporal facets of model performance. In some regions, where severe pollution episodes may be uncommon, and routine contaminant exposures from all other periods in a year are more of a concern, airshed managers could be more interested in simulation of air quality over annual, seasonal or even diurnal periods. To the author’s knowledge, no studies have been undertaken to identify the effect of different PBL parameterization schemes in the downscaling of meteorological variables for modeling the air quality in northwestern Canada. Since outputs from meteorological models are commonly fed into air quality models, ultimately weighing on airshed planning and policies, it is imperative to assess the reliability of downscaled meteorology prior to their use for current assessments and future projections. This chapter conducts high-resolution simulations of meteorological variables with theWRF mode (version 4.0), and the fifth major global reanalysis product of the European Centre 15 for Medium-range Weather Forecasts (ERA5), with a focus on evaluating performance of different PBL formulations. Its objectives are to quantify modeling uncertainties and to qualify mixing and dispersion capabilities of alternative physical schemes. The remainder of this manuscript is organized as follows. In section 2.2, the experimental set-up, including the PBL schemes being tested, nested domains, and WRF model configuration are described. Section 2.3 evaluates the uncertainty in modeled surface meteorology via comparisons with observations. Differences among simulations resulting from PBL scheme options are also analyzed. Section 2.4 discusses the implications of simulation similarities and differences for air quality modeling and concludes the chapter. 2.2 Methods 2.2.1 Description of the WRF model and PBL schemes WRF is a state-of-the-art atmospheric modeling system designed for both meteorological research and numerical weather prediction. The dynamical solver, known as the Advanced Research WRF (ARW) core, is based on the fully compressible, non-hydrostatic Euler equations with terrain-following Eta coordinates. WRF can run on a variety of computing platforms and suits a broad range of applications across scales ranging from tens of meters to thousands of kilometers (Skamarock et al. 2019). Turbulence closure schemes, also referred to as PBL schemes, are built-in to meteorological models to represent turbulent fluxes that are not explicitly resolved on the model grid. PBL processes are parameterized either through local closure or nonlocal closure models. 16 In the case of local closure, only those vertical levels that are adjacent to a given point directly affect variables at that point. Nonlocal closure schemes use multiple vertical levels to determine variables at a given point (Cohen et al. 2015). For meteorological downscaling with the WRF model, six such schemes are experimented. These schemes are the: Mellor–Yamada–Janjić (MYJ), Asymmetric Convective Model, version 2 (ACM2), Mellor–Yamada–Nakanishi–Niino level 3 (MYNN3), Shin–Hong (SH), University of Washington (UW), and Yonsei University (YSU). The YSU scheme is a first-order nonlocal scheme with explicit entrainment at the PBL top and a parabolic K-profile in an unstable mixed layer. It calculates PBL height from the surface and uses a threshold Richardson number of 0.25 for stable cases, and zero for unstable flow (Hong et al. 2006). The SH scheme is a nonlocal ’scale-aware’ formulation for vertical transport in a convective PBL. Vertical mixing in the stable PBL and free atmosphere follows YSU, in addition to diagnosed turbulent kinetic energy (TKE) and mixing length, but the explicit treatment of heat flux entrainment is replaced by grid sizedependent terms for nonlocal, and local transport components (Shin and Hong 2015). The ACM2 scheme is a hybrid scheme that features nonlocal transport from the lowest level to all other model layers alongside a local eddy-diffusion component, and an asymmetrical layer-by-layer downward transport. It calculates PBL height when the critical bulk Richardson number above the level of neutral buoyancy exceeds a value of 0.25. For stable or neutral flows, the scheme shuts off nonlocal transport and uses local closure (Pleim 2007). The MYJ scheme is a one-and-half-order local closure that includes a prognostic equation for TKE. The PBL height is the height at which the TKE reaches 0.2 m2 s-2 17 (Janjić 1994). The MYNN3 scheme is a second-order closure that expresses stability and mixing length based on the results of large eddy simulations rather than observations. It prescribes PBL height when TKE reaches 1.0 × 10-6 m2 s-2 (Nakanishi and Niino 2009). The UW scheme is a one-and-half-order local closure in which TKE is a diagnosed, rather than a prognostic variable. It uses moist-conserved variables with an explicit entrainment closure for convective layers. A threshold of 0.25 for the critical bulk Richardson number is used to determine PBL height in all stability cases (Bretherton and Park 2009). The above schemes are a mix of legacy and newer physics modules that have accompanied major updates to the WRF code. For instance, whereas the YSU, MYJ, and ACM2 schemes are quite common in the literature (e.g. Nielsen-Gammon et al. 2010, Zhang et al. 2013), the UW and SH schemes are fairly recent additions. Consequently, selected PBL schemes represent the state-of-the-art in options for parameterizing atmospheric processes over predominantly pristine (non-urbanized) areas. Each PBL formulation can be mapped to just one or more surface-layer schemes. The surface-layer schemes compute friction velocity and other exchange coefficients for estimating sensible and latent heat fluxes, and momentum flux from land surface models, and surface stress in the PBL schemes. In this study, the MYJ, MYNN3, and UW PBL schemes are coupled to the Eta surface-layer scheme (Janjić 2002), while the ACM, YSU, and SH PBL schemes are mapped to a modified MM5 surface-layer scheme (Jiménez et al. 2012). Both surface-layer schemes are based on similarity theory but the Eta similarity includes a parameterization of viscous sub-layer (Janjić 2002). 18 2.2.2 Domain configuration, model initialization and settings Computational domains assumed a telescopic nesting arrangement (Fig. 2.1). A grid ratio of 1:5 applied to successive parent domains on a Lambert Conformal projection was used to attain a horizontal resolution of 1 km for the TKV and surrounding areas (Table 2.1). Because there are several lakes in the valley, terrain data were interpolated from the Moderate Resolution Imaging Spectroradiometer (MODIS) International GeosphereBiosphere Programme (IGBP) 21-category land use and land cover fields with representation for lake bodies. This was to effectively distinguish between inland water bodies and seas, for appropriate processing and extrapolation of surface air temperatures over lakes. Atmospheric forcing data (for initial and boundary conditions) for the entire set-up was ERA5 (ECMWF 2019). Simulations up to a model top of 50 hPa were then performed for the year 2017. The 2017 meteorological year was selected because it best represents the climatological average (of wind speed and precipitation) for the valley from 2006 to the present. Appendix A has the details for this choice. For all simulations, physics settings (Table B.1 in Appendix B) were the Noah land surface model (Tewari et al. 2004), the Thompson scheme for microphysics (Thompson et al. 2008), the rapid radiative transfer model (Iacono et al. 2008) for longwave and shortwave radiation, and the horizontal Smagorinsky first-order closure (Talbot et al. 2012) for mixing terms. The Grell-Freitas ensemble cumulus physics scheme (Grell and Freitas 2014) used for the largest and intermediate domains was not required for the innermost (study) domain (Arakawa et al. 2011) due to its sufficiently high resolution. A one-way nesting (that is, the transfer of information from parent nests without feedback from child nests) was implemented. Since 19 meteorological data would be fed to an air quality model that required at least 10 days spin up period, WRF outputs were first generated for the period 00:00 UTC 20 December 2016– and 00:00 UTC 31 December 2016. Thereafter, WRF simulations with the ERA5 dataset were initialized monthly, with model spin-up of one day, and a further one day at the end of each month. The monthly overlap days, including for the period 20–31 December 2016 were discarded when merging hourly output files for the meteorological analyses. Figure 2.1 The Terrace– Kitimat Valley (TKV) and surroundings. Left: WRF nests for downscaling ERA5 data to study area. The immediate bounding nest for the TKV is d03 in red. Right: Enlargement of the d03 domain. Black outline specifies the valley channel wherein meteorological observation stations for model evaluation are indicated with black-filled markers. The red line is assumed centerline through the valley and the coastal channel 20 Table 2.1 Properties of the WRF model domains. Domains are roughly centered on the valley. Simulations for all three domains were into 39 vertical layers of varying thickness, with 12 layers in the lowest 2000 m. Geographic data of horizontal resolution similar to grid sizes of respective domains were deployed. Domain Horizontal x grid y grid Grid centre Grid centre Geographic (static) longitude data resolution resolution points points latitude d01 25 km 100 100 54.200 ◦ N 128.600 ◦ W 20-arc-minute d02 5 km 121 121 53.850 ◦ N 128.795 ◦ W 2-arc-minute d03 1 km 101 121 54.223 ◦ N 128.640 ◦ W 30-arc-second 2.2.3 Observational data Three surface meteorological variables were selected for the evaluations. These are 2meter air temperature (T2), specific humidity at 2 m (Q2) and 10m-wind (speed and direction). These variables were selected because they influence ambient air quality (Banks and Baldasano 2016, Dennis et al. 2010) and are monitored at various locations in the valley (Fig. 2.1, Table 2.2). Observation stations are operated and maintained by several organizations including Environment and Climate Change Canada, BC Ministry of Transportation and Infrastructure, and BC Ministry of Environment and Climate Change Strategy’s Air Quality Network. Station data can be downloaded from https://data.pacificclimate.org/ portal/pcds/map/. Because each network is designed to meet the specific need of its operating agency, there is variety in site characteristics and sensor equipment. For each parameter at any station, only records with at least 80 % yearlong completeness in hourly values are used for evaluation. 21 Table 2.2 Observation data locations for validation of surface meteorology variables for the TKV Station Latitude Longitude Elevation Station Description Surface variable 0.0 Maritime buoy Temperature, Wind 128.639 94.0 Residential area Temperature, Wind 128.538 220.0 Lakeside monitor (◦ N) (◦ W) (m) Nanakwa 53.830 128.830 Whitesail 54.067 Onion Lake 54.302 Temperature, Wind, Specific humidity Terrace 54.522 128.608 81.0 On school property Temperature, Wind, Specific humidity 2.2.4 Statistical evaluation indices For evaluations, values for model grid cells corresponding to ground locations of observations were retrieved. The fitness measures used in this work are the mean bias (MB), the root-mean-square error (RMSE), Pearson’s correlation coefficient (r) and modified index of agreement (IOA) according to Willmott et al. (2012). These indices are defined in Appendix C and were selected to reflect a mix of accuracy measurements for paired data. They also enable comparison to performance benchmarks (Table 2.3) for surface variables suggested by Emery et al. (2001). The benchmarks were developed over a limited duration in summertime, in an environment that is dissimilar to the TKV, therefore, may not be ideal for high-resolution, cross-seasonal simulations over complex terrain. MB as well as RMSE are in units of the parameter of interest and are negatively-oriented indices, 22 which means values closer to zero are better. For r and IOA however, scores range from -1 to 1, for which values approaching 1 indicate good performance. The identification of the most suitable PBL scheme is from a count of instances in which it is top-ranked, and also meets the performance benchmark. This is checked per statistical measure for all the stations. Table 2.3 Performance benchmarks (except for r) proposed by Emery et al. (2001) for surface meteorological variables in mesoscale models. Statistical measures per season, are on diurnal-hour means (0, 1, .., 23 ) for air temperature and wind speed, and on all hourly values for wind direction and specific humidity. Seasons are spring for the months of March to May, summer for June to August, autumn for September to November, and winter for December to February. RMSE is used in place of mean error for air temperature and specific humidity Benchmark Measure 2.3 Air temperature Specific humidity Wind direction Wind speed ±0.5 m s−1 MB ±0.5 K ±1 g kg−1 RMSE ≤2K ≤ 2 g kg−1 ≤ 2 m s −1 IOA ≥ 0.8 ≥ 0.6 ≥ 0.6 r ≥ 0.8 ≥ 0.8 ≥ 0.6 ±10 ◦ Results and analyses 2.3.1 Synoptic and mesoscale representations PBL schemes drive the vertical distribution of fluxes in the atmosphere and soundings of temperature, humidity and wind alongside modeled profiles from each PBL scheme 23 are examined qualitatively. All simulations reasonably retrace the above aerological variables for different times in winter and summer seasons at the Annette Island radiosonde station, Alaska, which is the upper air sounding location nearest the TKV (Fig. D.1 in Appendix D). This station is outside the 1-km grid, but within the 5-km grid. Temperature and dew point profiles modeled by the various schemes virtually match observations. Simulated dew point profiles for 0400 PST on a January day replicated a drying lower troposphere; coincident profiles of the YSU and SH schemes being drier than observations at 700 hPa height above the surface. Representation of the veering wind is also reasonable, given that specific days and times could have unusual atmospheric conditions. It needs emphasizing however, that the atmospheric environment around the Annette Island station, at least for the lowest 2000 m, could differ markedly from that of the TKV, since the station is west of the Coast Mountain ranges, completely exposed to Pacific Ocean air masses. Despite physiographical differences between the radiosonde location and stations in the TKV, the fact of general correspondence for upper air data attests to a reasonable representation of synoptic information. Figures 2.2 and 2.3 are plots of meteorological conditions at the surface for chosen summer (July 22) and winter (January 21) days, respectively, whose comparisons to observed seasonal averages of air temperature and wind are in Fig. E.1 of Appendix E. Specific times of the spatial plots are hours of diurnal temperature maximum and minimum over land, which are relevant for assessing mesoscale circulations such as land-sea breezes and mountain-slope winds. One verification of WRF simulation realism is the altitudinal decrease of daytime air temperature in summer, from around 18 °C on the valley floor to 24 about 4 °C over the mountains, which is consistent with the expectation of the valley to be warmer than the ridges (Fig. 2.2). Wind vectors are also more organized and aligned with the main and tributary valleys than outside them, indicating substantial wind adjustment to the underlying terrain and the efficacy of simulation at high horizontal resolution. Over the main valley itself, modeled winds are seasonally nuanced. In summer (Fig. 2.2), the wind is southerly, which is representative of the up-valley flow that occurs for ∼ 70 % of non-calms (wind speed ≥ 0.5 m s-1 ) observed at the Terrace station. This pattern is also dominant overnight (0200 PST) in all the PBL schemes, perhaps due to fewer nighttime hours in summer for reversal of wind direction, or from stronger synoptic forcing on this day, but terrain recognition by the WRF model appears satisfactory. Southerly wind vectors turn anticlockwise on approaching the Skeena River, resulting in a flow down the river channel, towards the coast. Some convergence of vectors towards watercourses that drain the valley, suggesting downslope winds in tributary valleys is also present in the simulations at 0200 PST. The afternoon hour (1400 PST) mainly indicates wind directions out of stream channels, or upslope. Compared to the nighttime period, the southerly flow in the valley is as would be expected, more intense due to enhancement by sea breezes. Wind vectors within the valley are straighter and better organized, with slightly more build-up on valley sidewalls for the MYNN3 and MYJ schemes. The winter season (Fig. 2.3) when the valley is snow-covered, differs markedly from summer. At night (0200 PST), the warmest parts are around the shoreline, and over major inland water surfaces. This is as expected since water bodies cool more slowly than land. Over the ridges that flank the valley, the MYNN3, MYJ, and UW PBL schemes depict 25 Figure 2.2 Ambient air temperature and wind vectors at the surface for a typical day in summer, at night (top) and during daylight (bottom) simulated by the various PBL schemes. 26 warmer temperatures than the ACM2 and YSU PBL schemes, probably due to differences in surface-layer schemes. By afternoon, the land areas have warmed such that both the valley and shore areas are mostly above 0 °C. Indeed, at 1400 PST, the northern half of the valley is warmer than the southern portion, with temperatures at mountain peaks similar (-9 to -6 °C) amongst PBL schemes. Wind vectors for the winter season are consistent with the dominance of northerly outflow at this time of a year; however, these are less organized than during summer. While the flow within the valley remains fairly distinct from winds over surrounding ridges, wind vectors in the valley are less organized than in summer suggesting stronger synoptic forcing in the winter season on this day. Once exiting the valley into the Douglas Channel, wind vectors are more aligned with the fjord, consistent with the expectation of less drag over water surfaces. 2.3.2 Performance evaluations for surface variables Ambient temperature Near-surface temperature (T2) is an important meteorological parameter as it affects the buoyancy of air pollutants and reaction rates of chemical species. Figure 2.4 shows diurnal sequences of hourly mean biases for summer and winter seasons. In summer, a warm bias occurs after sunrise, peaking at about 0800 PST, becoming a cool bias after noontime, and reaches a negative peak at sunset. The adaptation to convective transport by the nonlocal SH and YSU PBL schemes whose bias profiles are virtually the same, is evident in larger daytime bias amplitudes than those of the other schemes. At nighttime, YSU, SH and ACM2 have marginally colder biases than the others, leading to relatively cooler 27 Figure 2.3 Ambient air temperature and wind vectors at the surface for a typical day in winter, at night (top) and during daylight (bottom) simulated by the various PBL schemes. 28 temperatures over the course of a single day with these schemes. Figure 2.4 Hourly mean bias of T2 (modeled minus observation) over land for each PBL scheme in summer and winter seasons. Profiles above 0 K signifies warm bias. Dark grey plot backgrounds are nighttime periods. Table F.1 in Appendix F tabulates model performance across four seasons in a year. Going from the spring to winter period, T2 simulations change from negative biases to predominantly positive biases. Biases are less during the summer and winter seasons, compared to the spring and autumn seasons. Over land, diurnal MB ranged -2.5 to 1.3 K in spring, 29 0.6 to -0.2 K in summer, 0.5 to 4.1 K in autumn and -0.3 to 0.9 K in winter. Model errors for T2 are also smaller in summer and winter months; these periods having RMSE values that are ≤ 2 K at all three land stations. Temporal correlations are good, and the r benchmark is achieved in 100 % of evaluations from spring through autumn (> 0.9 in summer) and 67 % for winter. Few evaluations meet the IOA criteria; although the majority of values are within 0.60–0.79 across seasons. Results from all statistical measures indicate that T2 simulations over land are poorest in autumn. Biases and errors for the Nanakwa station in spring (MB range: -0.5 to 0.1 K, RMSE range: 1.2 to 1.4 K) and summer, (MB range: -1.0 to -0.4 K, RMSE range: 1.5 to 1.7 K) suggest reasonable accuracy of interpolated sea-surface temperatures in warmer months. In terms of individual PBL schemes, MYJ marginally outperforms UW with regard to accuracy and precision indices (MB, RMSE, and IOA), while MYNN3 performs best for pattern replication (r). The MYJ scheme also has several satisfactory correlation scores, thus more ranked by all four statistical measures. Specific humidity Another important meteorological variable from an air quality perspective, especially in controlling aerosol formation and transformation processes, is specific humidity (water vapor content). Performance statistics for specific humidity at 2-m height (Q2) for each PBL scheme per season, and at two stations is presented in Table F.2 of Appendix F. Across PBL schemes and relative to observations, the dominant pattern is a dry bias. Biases are larger at the Onion lake station which receives more precipitation than the Terrace station, and in summer when more evapotranspiration occurs, than in winter. These further attest 30 to the dynamical consistency of model outputs. At the Onion lake station, for instance, the MB range is -0.50 to -0.37 g kg-1 in winter, as against -0.68 to -0.19 g kg-1 in summer. Overall, model performance for specific humidity is remarkable. Ninety-four percent of evaluations for MB, 100 % for RMSE, 56 % for r and 65 % of IOA fall within respective cut-off values of satisfactory performance. With regard to individual PBL schemes, the MYNN3 scheme has the best fit for the Q2 simulation, based on the number of top-ranked and satisfactory evaluations for it. Indeed, because all schemes underestimate Q2, the MYNN3 scheme best ameliorates negative moisture biases. Wind Surface wind impacts the mixing and transport of air contaminants from source areas to receptors. Biases in modeled wind direction (WDIR10) at the different stations are presented in Table F.3 of Appendix F. The localized nature of wind over complex terrain is attested by biases at Onion Lake which is anticlockwise, unlike other stations. Noteworthy is that at the Nanakwa buoy, almost all evaluations meet the ± 10 ° criteria for wind direction bias. On land, model skill is reduced, arising from increased surface roughness, and wind flow being affected by multiple terrain elements. Only 8 % of evaluations for land stations are within the bias range for acceptable skill. However, 75 % of all evaluations for Terrace are ≤ 20 ° and all for Terrace and Whitesail are ≤ 30 °. The largest biases are at the Onion Lake station, in the central portion of the valley (up to ∼ -60 °). Using the same rule for assessing suitability, the MYNN3 and UW PBL schemes rank best for wind direction. 31 Wind speed uncertainty due to PBL scheme choices is also of interest and Fig. 2.5 shows diurnal sequences of hourly mean biases for 10 m-wind speed (WSPD10) in summer and winter. Differences between schemes rise to as much as 1.5 m s-1 in summer, but are smaller (∼ 1 m s-1 ) in winter. The valley atmosphere is more prone to decoupling the surface layer from layers aloft in winter due to increased stability, thereby lessening the sensitivity of surface winds to PBL parameterizations. Strong synoptic forcing from land-falling Pacific mid-latitude cyclones would account for why modeled WSPD10 has greater biases in winter than in summer. It is reasoned that the general overestimation of WSPD10 is due to WRF not sufficiently weakening winter storms as they transition from the smooth ocean to the rugged interior. Indeed, wind speed biases increase from the coast towards the northern part of the valley. Aside from complications of cross-flow of the Skeena River (see Fig. 2.1), and urban obstacles that attenuate real surface winds in Terrace, the general area to the north is dominated by mountainous topography. Wind speed overestimation has been reported in several experiments using WRF at similar horizontal grid resolutions over complex terrain (e.g. Avolio et al. 2017, Hariprasad et al. 2014, Zhang et al. 2013). The performance statistics for WSPD10 for all four seasons for each of the PBL schemes are in Table F.4 of Appendix F. Overestimations occur not only in winter and summer months but also through spring and autumn. Across land stations, diurnal MB ranged 0.4 to 2.7 m s-1 in spring, -0.2 to -2.5 m s-1 in summer, 1.3 to 3.0 m s-1 in autumn, and 1.3 to 3.7 m -1 in winter. Just 10 out of 72 evaluations (14 %) are within the MB benchmark for wind speed. Model errors for WSPD10 are also higher in autumn and winter months, 32 Figure 2.5 Hourly mean bias of WSPD10 (modeled minus observations) according to PBL schemes. A value of 0 m s-1 signifies no bias. Dark grey plot backgrounds are nighttime periods. although the proportion of evaluations within the RMSE criteria is much improved (68 %). Favorable RMSE and r scores are in contrast with negative 1OA values. Whereas 64 % of evaluations for r are ≥ 0.6, only 11 % — all for Onion Lake— meet the IOA benchmark. Thus, compared to air temperature and humidity, WSPD10 fits poorly with the benchmarks, especially at Terrace. 33 2.3.3 Spatio-differential examinations of surface meteorological fields Because diurnal surface heating and cooling impact atmospheric structure, spatial differences in model outputs are examined for the relative effect of individual WRF PBL schemes on meteorological properties at nighttime and daylight periods. The seasonal attributes of PBL scheme adjustment to local weather and evolving capacities for mixing of surface fluxes are also qualified. Nighttime Some turbulence closure approaches are conditionally operable. The ACM2 scheme deactivates nonlocal transport for stable conditions, which would prevail at night. Of interest therefore, is how meteorological quantities for periods when only the local closure component of this scheme operates, compares to those from the fully nonlocal YSU scheme, given that both PBL schemes are mapped to the same surface layer model. Figure 2.6a shows that over the valley portion of the domain, the skin temperature (SKINT), T2, and WSPD10 from the ACM2 scheme are greater than those of the YSU scheme. At higher elevations outside the valley channel, the topography is less confined and the YSU scheme is more capable of moving cold air from the surface to layers aloft, and downward momentum transport. The relative weakening of nonlocal transport over the valley therefore derives from the confinement of the land surface by steep mountains, and is also due to nighttime stable conditions that restrain vertical transport through the entire PBL. In the corresponding wintertime, when the atmosphere is more stable, the temperature difference over the valley between the YSU scheme and the ACM2 scheme is increased (∼ 0.5 34 °C) but less widespread, since the severest restriction to nonlocal flux exchange would be at the deepest of cold air pools. Since the expression of Q2 as a mixing ratio is analogous to that of an air pollutant that may be discharged at the ground level, its comparison across schemes could be useful proxies for contaminant accumulation and dispersion. Greater Q2 with the YSU scheme over most parts of the study domain may be linked to the relative capacity of nonlocal closure to facilitate vertical transfers between model layers. But the correspondence of ACM2 /YSU paired differences for PBL heights (PBLH) and Q2 is also noteworthy. Relative to outputs from the YSU scheme, the PBLH simulated by the ACM2 scheme is thicker over water bodies, which are the same areas where its Q2 values are moister. In the YSU scheme, vertical diffusion is constrained to the diagnosed PBL depth. Thus, it is less diffusive of Q2 over the land areas. Over water, Q2 from the YSU scheme is subject to a steeper vertical gradient that enables faster transport of water vapor from the bottom of the PBL to overlying model layers. The constraining of nighttime mixing by steep topography is less with local turbulence closure. However various local mixing approaches exist whose expressions may also differ with alternative surface-layer parameterizations. Figure 2.6b compares the ACM2 scheme nighttime outputs, to those of the local MYJ scheme that is mapped to the Eta surface-layer model. Unlike the MYJ scheme, the local closure implementation in the ACM2 scheme ignores TKE transport. Hence in the winter season when temperature inversions are frequent, along-valley PBLH simulated by the ACM2 scheme is shallower than that of the MYJ scheme. However, that up to1.0 °C difference for T2 over the TKV is 35 modeled, suggests surface-layer scheme influence. Over the valley, MYJ outputs warmer T2 than ACM2, but cooler skin temperatures (SKINT), hinting at greater heat flux retention with the Eta surface-layer scheme than with the MM5 surface-layer scheme. Unlike the MM5 surface-layer scheme, the Eta surface-layer scheme considers thermal roughness length (Janjić 2002), apparently contributing to warmer ambient temperatures with the MYJ PBL scheme. Differences in friction velocities (not shown) were similar to patterns for wind speeds, for which the MYJ scheme values over the valley and coastal channels are less. Friction velocity is a fraction of wind speed (Camuffo 2014). The friction velocity limit in previous versions of the MM5 surface-layer scheme ( 0.1 m s-1 ) was considered as preventing undesirable effects such as excessive ground cooling (Jiménez et al. 2012). The revised MM5 surface-layer scheme as used in the present study, permits friction velocities as low as 0.01 m s-1 (Jiménez et al. 2012), while in the Eta surface-layer scheme, a correction is applied so that only non-zero values are obtained (Janjić, 2002). Consequently, lower friction velocity over the valley and coastal channels with the MYJ scheme possibly derives from its coupling with the Eta surface-layer scheme, which in turn, could offset part of the warming effect of thermal roughness length. Cooler SKINT of the MYJ output relative to that of the ACM2 simulation over the valley channel may have resulted from such differences in friction velocities. For other model runs with the Eta surface-layer scheme (Figs. 2.6c-d), differences among PBL schemes for simulated variables are less than in Fig. 2.6b, where both the surfacelayer scheme and PBL schemes are varied. Nonetheless, the differential subtleties are worth examining, given that PBL scheme development, testing and revising, has primar- 36 ily centered on weather prediction, rather than on air pollution modeling. Warmer skin temperature for MYNN3 compared to MYJ translate to a higher T2, but do not result in greater water vapor retention over the entire grid. MYNN3’s summertime Q2 over the southern portion of the domain is drier than that of the MYJ scheme. But this portion roughly coincides with the parts where the height of MYNN3’s PBL surpasses that of the MYJ scheme (also Fig. G.1 in Appendix G), suggesting greater mixing with the MYNN3 scheme. This seems to impact nighttime WSPD10 as well, with greater (lesser) values for the MYNN3 scheme coastward (landward). In addition to implementing local eddy diffusion, the MYNN3 scheme, unlike the MYJ scheme, has a nonlocal mixing component. The suppression of nonlocal vertical exchange in the MYNN3 scheme relative to the MYJ scheme, is therefore less than for the YSU scheme in relation to the ACM2 scheme in Fig. 2.6a. There is a seasonal signal with this behavior. In summer, modeled onshore winds generate instabilities over considerable distances upon which nonlocal mixing is enhanced. In wintertime, when valley air is colder and drier, and simulated circulations are northerly, MYNN3’s nonlocal mixing is restrained: the mixing is more diffusive than in the MYJ scheme only over the Douglas channel that is exposed to warmer, maritime air. The UW scheme, designed for simulation of marine stratocumulus-capped boundary layers, neglects TKE storage, and assumes no TKE transport occurs during stable conditions (Bretherton and Park 2009), hence it simulates the lowest PBL heights and cloud bases that are nearest to the ground (Figure G.1 in Appendix G). Some modeling experiments in mountainous regions (e.g. Balzarini et al. 2014) have also reported lower PBL heights 37 Figure 2.6 Spatial differences of averaged skin temperature (SKINT), ambient temperature (T2), surface wind speed (WSPD10), ambient moisture content (Q2) and PBL heights at 0200 PST in summer and winter. MYNN3 – MYJ means values for MYJ subtracted from MYNN3, etc with the UW scheme. Its output of smaller Q2 than that of the MYJ scheme (not shown) suggests a relatively rapid moisture loss mechanism. It is worth stating that the UW 38 PBL scheme’s verification and integration into the WRF model noted the diffusion of humidity and chemical species from the lowest model grid level, rather than their mixing across layers (Bretherton and Park 2009). Because air pollutants are also mass fluxes, less nighttime ambient moisture with the UW scheme relative to the MY-schemes suggests pollutant concentrations could be lower with the UW scheme. Daytime PBL schemes adjust differently to enhanced convection in the daytime, also impacting meteorological fields over complex terrain. Comparisons between the ACM2 and YSU schemes in summer (Fig. 2.7a) indicate that for the most part, the ACM2 scheme produces a higher PBL over land areas, in reverse of the nighttime pattern. Unlike single estimation of the PBL top in the YSU scheme, the PBLH in the ACM2 scheme is calculated from separate diagnosis of convective and entrainment layers (Pleim 2007), partly explaining PBLH differences. Convective regimes activate the nonlocal transport component of ACM2, thereby providing a hybrid vertical flux transfer mechanism, as opposed to the YSU scheme that is nonlocal. Over valley areas adjacent to the marine channel, incursions of maritime air may diminish the development of convective regimes more than at positions farther inland, hence small portions where the ACM2 scheme’s Q2 output surpass those of the YSU scheme. WSPD10 differences along the Douglas Channel show the ACM2 output is windier, also contributing to moister inflow. But for the most part, ACM2 generates less Q2 in the daytime of summer (see also Fig. 2.8), which suggests stronger updrafts by the ACM2 than by the YSU scheme in this period. Therefore, the 39 ACM2 scheme could produce less pollutant concentration at the surface than the YSU scheme, especially in the drier, northern part of the valley. In the cold winter season (Figs. 2.7a, 2.8), the daytime Q2 difference is negligible (< 0.1 g kg −1 ) and the change of water vapor with height is comparable between the two PBL schemes. This is because the ACM2-predicted PBL heights are mainly lower than those of the YSU scheme, and its nonlocal transport component is inactive in stable/neutral conditions that dominate in winter. Figure 2.7b compares functionally-similar PBL schemes for daytime mixing but mapped to different surface-layer schemes. The MYNN3 scheme, with relatively positive incoming radiation over the valley in summer, simulates lower SKINT than the ACM2 scheme. However, over the mountains, T2 differences, even in wintertime, are more positive for the MYNN3 scheme than over the corresponding areas in the SKINT plot. This again could be due to greater sensible heat flux from the Eta surface-layer scheme. Apart from differences in surface-layer formulation, dissimilarity in the implementation of hybrid mixing methods would also account for differences in simulated quantities. The K-theory based ACM2 scheme uses first-order statistics to approximate higher moment terms, while the MYNN3 scheme is a TKE-based second-order closure. The PBLH diagnosed by the ACM2 scheme is predominantly thinner (thicker) in summer (winter) than the MYNN3. Adjustment of vertical gradients of momentum and moisture fluxes with PBLH changes in the ACM2 would also affect WSPD10 and Q2, respectively. The MYNN3 scheme mostly outputs greater Q2 and WSPD10 than the ACM2 scheme but the extent to which PBL parameterization alone accounts for this outcome is confounded by 40 the use of alternative surface-layer schemes. Simulation differences among the TKE-based PBL schemes (Figs. 2.7c-d) show the influence of local and nonlocal closures with the Eta surface-layer scheme. PBLH is greater with MYJ than with MYNN3 except near the coast. As previously stated, the daytime intrusion of marine air counteracts growing instability arising from solar heating. For this reason, higher PBLH with the MYNN3 scheme relative to that of the MYJ scheme does not extend as far inland as it does at night (Figure G.1 in Appendix G). But instead of a moisture profile that is more vertically homogeneous than those of the MYJ and UW schemes, the MYNN3 scheme indicates less mixing (Fig. 2.8) due to how it implements nonlocal transport. The MYNN3 scheme comprises a partial-condensation model for convective activity, in which the buoyancy effects of subgrid-scale clouds are considered using counter-gradient terms (Nakanishi and Niino 2009). The counter-gradient flux facility in the MYNN3 scheme thus explains why it has greater (lesser) Q2 (WSPD10) values than the outputs from the MYJ and UW schemes. Recent revisions to mixing-length formulations in the MYNN3 scheme for which the current code further decreases the downward mixing of momentum (Olson et al. 2019), would also contribute to its generation of WSPD10 and Q2 values that are less biased from station observations than those of the UW and MYJ schemes. Because of increased linking of the MYNN3 scheme to cloud processes and to other physics modules, such as radiation schemes (beginning from WRF version 3.8), feedbacks from simulated boundary layer clouds are probably more influential on surface meteorological variables than with other PBL schemes. For instance, the UW scheme outputs greater cloud coverage than 41 Figure 2.7 Spatial differences of averaged skin temperature (SKINT), ambient temperature (T2), radiation reaching the surface (INRAD), surface wind speed (WSPD10), ambient moisture content (Q2) and PBL heights at 1400 PST in summer and winter. MYNN3 – MYJ means values for MYJ subtracted from MYNN3, etc. the MYNN3 scheme (Figure G.1 in Appendix G), which ought to result in less radiation reaching the surface with the UW scheme. Instead, incoming solar radiation is smaller with the MYNN3 scheme, consequently its colder T2. The consideration of moist convection by the MYNN3 scheme is also evident in it having the biggest gradient for daytime 42 potential temperature (Fig. 2.9). It is reasoned that the facilitation of cloud condensation in the MYNN3 scheme, enabling latent release and warming of the air aloft, alongside less solar radiation to the surface and cooler near-surface temperature, causes a stronger temperature gradient. The least vertically-homogenized moisture gradient within the lowest 500 m for the MYNN3 scheme, may partly be due to this stability-enforcing feedback. In the wintertime when absolute water vapor content is less, and convective activity is reduced, MYNN3’s Q2 surplus compared to those of MYJ and UW in the valley, is diminished or in deficit. The vertical moisture gradient at this time is slightly steeper for the MYJ scheme than for the MYNN3 scheme. The above analysis point to the seasonal modulation of MYNN3’s nonlocal mixing. 2.4 Discussion and Conclusion Performance scores of the various PBL schemes for surface meteorological parameters are generally within, or close to values that are found in literature, either as NWP experiments (García-Díez et al. 2013, Hu et al. 2010, Zhang et al. 2013) or in the validation of meteorological inputs to air quality models in mountainous regions (Cheng et al. 2012, Hedley and Singleton 1997, Mues et al. 2018). It is not unusual for performance to vary between models, model versions, grid resolutions, episode durations, and performance is also subject to operating environments and needs. As an example, the evaluation benchmarks in Emery et al. (2001) were based on MM5 model application at 4-km grids to two ozone episodes, each lasting a week, in Texas, US. Further, there usually is a minimum amount of scatter or representativeness error that is inherent in meteorological downscal43 Figure 2.8 Averaged vertical profiles of simulations for mean water vapour mixing ratio in the lowest 1000 m at 0200 and 1400 PST in summer and winter ing and which cannot be removed. Rife et al. (2004) estimated representativeness error in a model grid-box size of 1.33 km2 to be about 1 m s-1 for near-surface wind speed in a well-mixed boundary layer over complex terrain. This type of error would increase with a smaller model grid size such as in the present study. Therefore, error and bias estimations previously presented reflect the status of the WRF model (version 4.0) and ERA5 to simulate surface variables in the TKV without observation nudging. Due to atmospheric circulations constrained by steep topography, the performances of 44 Figure 2.9 Averaged vertical profiles of simulations for mean potential temperatures at 0200 and 1400 PST in summer and winter. Computed gradients are for the lowest 1500 m the PBL schemes for wind direction were reasonably comparable. Therefore, the choice of one scheme over another may not be critical to representing the relative contributions of wind sectors to pollutant concentrations at specific locations. Performances for surface air temperature and wind speeds, on the other hand, were more varied. Overpredicted wind speeds in particular could cause an underestimation of ambient pollutant concen- 45 trations in areas near major emissions sources. Indeed, the finding that PBL schemes were able to simulate surface wind with sufficient directional accuracy, yet poorly represent the observed speed, summarizes the challenge of validating wind fields at high spatial resolution over heterogeneous terrain for input to coupled numerical models. As was noted previously, dynamical downscaling aligned gridded coarse meteorology to the valley and fjord channels, hence the general correspondence of modeled wind direction with observations, while ambient wind speed biases reflected uncertainties that can arise from smoothening of rugged relief in the simulations. Model terrain height errors were a few to several dozen meters, and terrain data quality and resolutions may have affected individual simulations differently. It was noticed however that performances for the YSU and SH schemes were similar in nearly all aspects, hence differences in concentrations of air pollutants that may arise from selecting either scheme would probably be negligible. Both schemes are founded on nonlocal Kprofile assumptions, but the SH PBL scheme additionally addresses turbulent transport in the ’gray zone’ (Wyngaard 2004), which is the range of grid spacing below which parameterizations have been developed, but above which the relevant processes are entirely resolved. Since simulations at 1-km horizontal grid spacing as used in the present study may be coarser than gray-zone resolution, the grid-size dependency of sub-grid scale, nonlocal heat and momentum transport that is unique to the SH scheme, was probably not optimized. In a real-case verification of the SH scheme by Shin and Hong (2015), the simulation of convective rolls was structurally more distinct from those of the YSU scheme at less than 1-km grid spacing. Because of the 0.281 ° spatial resolution of the 46 ERA5 forcing dataset, implementing a smaller grid size for differential modeling over the TKV was not done since that would have resulted in more computing costs, and sharp elevation gradients causing numerical instabilities in the simulations. Further, convective periods in the TKV are restricted to daytime in the summer, and involve moist turbulence which is not represented by the SH scheme. Distinctions in model performance generally evolved with season changes. In summer when Pacific frontal passages are rare and local thermally-driven circulations are enhanced, the diurnal signatures of individual PBL schemes were fairly unique. Particularly for air pollutants that peak in the warm season such as ozone, the selection of one PBL scheme over another may be significant. In winter, synoptic forcing is stronger, the daylight period is shorter and diurnal variability is weaker than in summer. More extended decoupling of the atmosphere in wintertime diminished the influence of model configurations on near-surface variables, and all schemes performed poorly in this period. But because atmospheric phenomena such as temperature inversions are not always surfacebased, and in the absence of radiosonde measurements, ensemble projections of the air quality in the cold season rather than the output from just one PBL scheme may be better. The UW PBL scheme that consistently produced the shallowest boundary-layer heights may need further investigation to know how realistic such representations are for air quality parameters in the region. In essence, PBL schemes are highly integrated with other physics and dynamics options. Some of the PBL schemes that were tested in this work can be mapped to other surfacelayer models in WRF. Specifically, the ACM2 PBL scheme can also be mapped to the 47 Pleim-Xiu surface-layer scheme, while the MM5 surface-layer scheme can be used for both the MYNN3 and UW PBL schemes. As an example, compared to the Eta surfacelayer scheme, MYNN3 mapped to the MM5 surface-layer scheme produced larger negative biases for temperature, specific humidity and wind speed (Fig. 2.10). Where a turbulence closure scheme can link to more than one surface-layer model option, the practice was to map TKE-based PBL schemes to the Eta surface-layer scheme, and first-order closure schemes to the MM5 surface-layer scheme. Figure 2.10 partially strengthens this practice. This study refrained from comparing modeled values to domain-wide averaged observation data. One shortcoming in several past studies was limited objective verification of target meteorological parameters through time and non-clarity as to what qualifies as good performance of mesoscale models in regions of strong surface heterogeneity. The present approach has been to assess performance at discrete locations and to supplement conventional metrics such as MB, RMSE and r scores with a modified IOA (bounded between -1 and 1) that evaluates the strength of model predicted variability relative to observed variability. As an example, summertime RMSE and r values for wind speed at the Onion Lake and Whitesail stations were comparable and within performance benchmarks. However, while IOA values for Onion Lake were good, IOA for Whitesail was poor and attributable to errors (overestimation) that are much larger than the variability in observations. This is in spite of the fact that the Whitesail station is located nearer the shoreline and should ideally match better with strong breezes that were simulated. It is important to consider such error disparities in downscaled meteorology over complex 48 coastal areas when simulating long-term chemical concentrations. Figure 2.10 Summer and winter hourly mean bias plots of T2, Q2 and WSPD10 for Eta similarity (solid lines), and revised MM5 similarity (dashed lines) surfacelayer schemes respectively, mapped to MYNN3 PBL scheme. 49 Chapter Three Evaluation of CMAQ modeling sensitivity to planetary boundary layer parameterizations for gaseous and particulate pollutants over a fjord valley (Also published at https://doi.org/10.1016/j.atmosenv.2020.117607 in Atmos. Environ.) Abstract Three-dimensional chemical transport models are useful for spatial and temporal analyses of outdoor air quality. However, the suitability of boundary-layer parameterizations for air pollution modeling over deep, coastal valleys has seldom been tested. An evaluation of the Community Multiscale Air Quality (CMAQ) model performance for five planetary boundary-layer schemes (PBL) with the Weather Research and Forecasting (WRF) meteorological driver was conducted at 1-km horizontal resolution for fine particulate 50 matter (PM2.5 ), sulfur dioxide (SO2 ) and nitrogen dioxide (NO2 ) over the Terrace-Kitimat Valley. The top-ranked schemes were Mellor-Yamada-Nakanishi-Niino Level 3 (MYNN3) for PM2.5 and Mellor-Yamada-Janjic for NO2 . Both schemes ranked high for absolute SO2 levels, but the MYJ and Asymmetric Convective Model, version 2 (ACM2) schemes qualitatively emulated peak summertime diurnal concentrations in the near field of elevated point sources. Greater nighttime SO2 concentrations with MYNN3 and Yonsei University PBL schemes, in less agreement with station monitoring 8 km downwind of emissions from tall stacks, suggested sustained pollutant mixing and downward transport within the nocturnal boundary layer. Consequently, for these two schemes with representations of nonlocal mass flux transfers between model layers, inland penetrations of pollutant plumes were farther than those of ACM2, MYJ, and University of Washington schemes. For NO2 and PM2.5 that mainly discharged passively from fugitive, ground-level sources, hence are less accurately quantified than SO2 emissions, the fully local MYJ, and semilocal MYNN3 PBL schemes more reasonably reproduced peak season concentrations than other schemes. It is concluded that for air pollution modeling in rugged, remote areas, the mode of pollutant emissions is important for the choice of a PBL scheme. PM2.5 was consistently underestimated by the various PBL schemes, and aspects for improving CMAQ simulations for a complex environment are discussed. 3.1 Introduction Advanced computations coupled to numerical weather prediction (NWP) models involve several physics modules for convection, cloud formation, boundary-layer turbulence, ra51 diation and land-surface processes, that control circulations at various spatial scales, and are important for the evolution of near-surface meteorological parameters. Of these, the representation of subgrid-scale turbulence, which determines the stability of the lowest portion of the atmosphere, is essential to environmental applications, such as simulating air quality. Especially for areas with sparse pollution observational networks, the choice of a planetary boundary layer (PBL) scheme, whose optimality may depend on weather regimes and phenomena, can affect the prediction accuracy of a chemical transport model. Numerical simulations with horizontal grid of several kilometers are often insufficient for addressing more localized air quality concerns, hence many have researched the influence of PBL parameterizations with finer (1–2 km) horizontal meshes. Li et al. (2019) evaluated two different PBL schemes in the Weather Research and Forecasting model coupled with Chemistry (WRF-Chem) and a Large-Eddy Simulation (LES) PBL model over the Baltimore-Washington region of the US. They found that the LES model afforded more realistic reactive chemistry for secondary pollutants formation. Pérez et al. (2006) analyzed air chemistry predictions from three PBL schemes over Barcelona, Spain and related daily maximum pollutant concentrations to mixing heights, temperature and wind speeds generated by each scheme. Parra (2018) assessed six PBL schemes for Cuenca, Ecuador and found that two schemes were satisfactory for short-term air quality observations. However, the majority focus of past studies has been for short durations and in areas with high background pollution, whose conclusions may not be valid for pristine regions. 52 The focus of this chapter is on the application of the Community Multiscale Air Quality (CMAQ) model (Appel et al. 2018) over the TKV. This model will be run at a spatial resolution not often used (1 km), and for pollutants that are routinely monitored within the airshed namely fine particulate matter (PM2.5 ), sulfur dioxide (SO2 ) and nitrogen dioxide (NO2 ). The objectives of this study are to (i) evaluate the overall performance of the CMAQ model across various monitoring stations in the TKV, (ii) identify the most suitable PBL scheme(s) for simulating air pollutants, (iii) examine the extent to which major air pollutants are sensitive to PBL parameterizations, (iv) suggest where improvements to chemistry-transport modeling may be needed. The remainder of this chapter is presented in three sections. The modeling approach, measurement data, and performance evaluation indices are described in the next section. Section 3.3, conducts a comparative analysis of results. Section 3.4 discusses the spatio-temporal context of biases and uncertainties relating to changes in PBL schemes and concludes the chapter. 3.2 Methods and data 3.2.1 PBL schemes and experiments set-up The downscaled meteorology from five PBL parameterizations schemes namely: MYNN3, MYJ, UW, YSU, and ACM2 in Chapter 2 were deployed. The Meteorology-Chemistry Interface Processor (Otte and Pleim 2010), version 4.3 was used to create two emissions grids from the WRF outputs (Fig. 3.1, Table 3.1). The Sparse Matrix Operator Kernel Emissions (SMOKE) modeling system, version 2.6 (Carolina Environmental Program 2009) 53 was used to prepare anthropogenic emissions for input to CMAQ. For each PBL scheme, emissions were modeled for the 5 km and 1 km grids. The British Columbia portion of Canada-wide emissions inventory, updated to 2017 for point emissions in the TKV was used. Stack emissions were processed as elevated point sources. Unique emissions in point source format, such as from aircraft at airports, were treated as ground sources. Spatial allocation of area (e.g. agricultural, road and construction dust) and mobile (onroad vehicles and off-road equipment) emissions with SMOKE was implemented with 124 gridding factors (surrogates). Among the gridding surrogates were population and housing, road classes and length, waterways, and mineral extraction, all resulting in allotment of provincial emissions to grid cells that intersect respective domains. Temporal splitting of area emissions with SMOKE into hourly time steps required by CMAQ was done based on local knowledge of activity patterns at diurnal, weekly and monthly intervals. Other important files necessitated by the long (1-year) simulation duration and geographical settings of our study, such as a seasonal switch file (for biogenic emissions), a land-sea mask (for sea salt), and a land-use grid (for agricultural dust emissions), were generated with either SMOKE or spatial allocator tools. The CMAQ chemical-transport model, version 5.2 (Appel et al. 2018) was run from 00 UTC 20 December, 2016, to 00 UTC 2 January, 2018. To implement a nesting mode with CMAQ runs (see Fig. 3.1), simulations were performed in two sets. The first set was for generating boundary concentrations for each PBL scheme from their respective coarse (5-km) emission grids. Thus, boundary conditions over the local (TKV) domain varied for each hour, as opposed to directly using clean, invariant North American profiles that are packaged with CMAQ. For the second 54 Figure 3.1 Telescopic grid nesting set-up for WRF-CMAQ modeling over the TKV. Left: 2-nests level of WRF domains set-up. The red box is the TKV’s meteorological domain. Right: 1-nest level of SMOKE emissions/CMAQ domains set-up. DOM_BC (in orange), smaller than the white bounds of d02 on the right, specifies the areal extent of emissions from which boundary concentrations to TKV’s air quality modeling domain (in black) were derived. The red box remains TKV’s meteorological domain. set of simulations which was for the 1-km grid, the first 12 days were model spin-ups and discarded, retaining only hours in 2017. All model runs used the Aero6 aerosol module (Appel et al. 2018) and carbon bond 05 for chemical speciation (Yarwood et al. 2005). All simulations were off-line (that is without feedback from CMAQ to meteorological fields). However, natural emissions from vegetation (biogenic), sea salt and wind-blown dust were calculated and treated in-line by CMAQ. Although the same number of layers as the input meteorology (Table B.2 in Appendix B) was used in CMAQ modeling, outputs were processed for only the layer nearest the surface since no upper-air pollution moni- 55 Table 3.1 Meteorological and emissions domain sizes and attributes. Emissions domains, with one level of nesting only, are the same sizes as CMAQ domains Domains (dimensions) Meteorology SMOKE Emissions/ CMAQ Concentrations Parent (W-E × N-S) 100 ×100 - Parent grid center 54.200 ◦ N, 128.600 ◦ W - Parent grid resolution 25 km - 1st nested (W-E × N-S) 121 × 121 40 × 60 1st nested grid center 53.850 ◦ N, 128.795 ◦ W 54.200 ◦ N, 128.600 ◦ W 1st nested grid resolution 5 km 5 km 2nd nested (W-E × N-S) 101 × 121 36 × 106 2nd nested grid center 54.223 ◦ N, 128.640 ◦ W 54.200 ◦ N, 128.600 ◦ W 2nd nested grid resolution 1 km 1 km toring exists in the study area. 3.2.2 Air quality monitoring data The most challenging air quality problems involve complex multi-pollutant interactions and coupling between atmospheric chemistry and dynamics. The performances of the PBL schemes for PM2.5 , SO2 , and NO2 simulations are of interest. These pollutants are directly released from various sources, but some portion of PM2.5 can be formed from reactions of other pollutants. The pollution monitoring network in the valley consists of five unevenly-spaced stations (Fig. 1.1, Table 3.2) categorized as urban, semi-urban, industrial and rural. Most stations do not monitor all three pollutants, for example, NO2 56 Table 3.2 Air quality monitoring stations in the Terrace-Kitimat Valley. All are ground-based and within the valley. Station Latitude Longitude Elevation Station Description Air pollutant 0.0 Rural SO2 128.702 11.0 Industrial (Kitimat) PM2.5 54.054 128.671 18.0 Semi-urban (Kitimat) PM2.5 Whitesail 54.067 128.639 94.0 Semi-urban (Kitimat) SO2 Terrace 54.522 128.608 81.0 Urban PM2.5 , SO2 , NO2 ◦N ◦W (m) Kitamaat Village 53.830 128.830 Haul Road 54.029 Riverlodge is only measured at Terrace. This station is operated by the BC Ministry of Environment and Climate Change Strategy (BCMOECS). The others are operated and maintained by an industrial permittee, reporting to the BCMOECS air quality network database (https: //envistaweb.env.gov.bc.ca/DynamicTable2.aspx?G_ID=327). Sensors in use are those recognized as meeting the United States Environmental Protection Agency’s Class III FEM monitors (USEPA 2016) standards. For each pollutant, only stations with at least 80 % completeness in valid hourly values are used for model evaluation. Low-level concentrations were not excluded from measurement data. 3.2.3 Statistical evaluation indices Statistical measures are used to quantitatively assess the predictive skill of a model. Modeled values for the grid cells coincident with monitoring stations were retrieved. Several authors (e.g. Chang and Hanna 2004, Emery et al. 2017) recommend the use of multiple 57 fitness measures for air quality model evaluations as no single measure is universally applicable to all conditions. The fitness measures used in this work reflect a mix of accuracy measurements for paired data and are listed and defined in Table 3.3. Performance evaluations are conducted by pairwise comparison of the diurnal profiles of measured, and modeled concentrations, for peak seasons. Where they exist, model performance criteria (Boylan and Russell 2006, Emery et al. 2017) are also used to rank statistical scores. An important caveat is that existing benchmarks are mostly from studies conducted with coarser grid resolutions, over flatter, more homogeneous terrains than the TKV area (see Emery et al. 2017), hence, may not be fitting for the present study. Table 3.3 Fitness measures based on paired hourly means (0, 1, .., 23). Formulas are defined in Appendix C. FAC2 and NMB criteria are from Chang and Hanna (2004). Performance criteria for PM2.5 are based on Boylan and Russell (2006) and Emery et al. (2017). Performance Measure Abbreviation Ideal Recommended FAC2 All ≥ 50% Mean Bias MB 0 µg m−3 / ppb Normalized Mean Bias NMB 0 ±0.3 Mean Fractional Bias MFB 0 ±0.6 (for PM2.5 ) NSD 1 Number of predictions within a factor of 2 of observations Standard deviation of modeled values divided by those of observations 58 3.3 Results and analyses 3.3.1 Ambient air quality Ambient concentrations in pollutant peak seasons are used for analyzing CMAQ simulations. For each station and pollutant, peak season was defined as the 3-month seasonperiod in which the mean hourly pollutant concentration is highest. At all three stations (Table 3.4) the peak season for ambient SO2 is in summer (June to August). The peak for NO2 in Terrace is in winter (December to February). For PM2.5 , the peak season for Haul Road and Terrace stations are in autumn (September to November) while that for Riverlodge is in winter. Table 3.4 Seasonal mean concentrations of air pollutants at model validation stations. Peak seasons are those with the highest seasonal mean of a pollutant (in italics). Pollutant PM2.5 (µg m−3 ) SO2 (ppb) NO2 (ppb) Station Spring Summer Autumn Winter Annual Haul Road 2.9 4.0 4.4 3.3 3.6 Riverlodge 3.6 3.8 4.6 5.4 4.4 Terrace 4.0 4.6 10.4 6.9 6.4 Kitamaat 0.31 0.31 0.29 0.27 0.29 Whitesail 0.47 0.61 0.28 0.27 0.41 Terrace 0.51 0.72 0.27 0.32 0.45 Terrace 2.0 1.1 2.5 4.7 2.7 Meteorological and local factors account for differing PM2.5 peak seasons at monitor loca59 tions. Winds are weaker in autumn than during summer and winter seasons. Especially at the Terrace station, where the frequency of nighttime calms is 30 % (Fig. 3.2), autumn PM2.5 average is more than twice the summertime value. Northerly wind sectors are slightly more frequent than southerly winds in autumn, hence the northerly PM2.5 sectors are the more prominent at Terrace, but at the Haul Road station in Kitimat, southerly sectors dominate PM2.5 levels. Actual PM2.5 concentrations (Fig. 3.3) are also similar among pollutant sectors at Terrace, unlike at Haul Road where the northerly sectors are much cleaner than the southerly sectors. Urban releases (e.g. from road traffic, domestic emissions, commercial outlets) are to all directions around the Terrace monitor, consequently similar qualitative influence. The Haul Road station is located in the industrial area of Kitimat, just north of a large aluminum smelter, but its precinct is composed of natural vegetation (uninhabited). Thus, the anthropogenic signal is more conspicuous for the Haul Road station, even though ambient PM2.5 at that location is much less than at the Terrace station. Advection of precursor smelter emissions by predominantly southerly winds in the warm season would account for summertime SO2 peak at Whitesail and Terrace stations. The primary source of SO2 in the Kitimat area and the TKV as a whole is the smelter site which emitted 27 tonnes of SO2 per day in 2017 (ECCC 2020). The strong south-westerly component at the Kitamaat Village station suggests intermittent, southward deflection of SO2 plumes, enough to produce mean concentrations that are comparable to other seasons when it is less upwind of the smelter source. Kitamaat Village however, is situated adjacent to maritime waters and stronger breezes there than inland, disperse pollutants 60 better. Figure 3.2 Wind roses as wind frequency proportions at Terrace and Whitesail stations for summer, autumn and winter periods. Radial increments are 15 %, with the largest radial at 45 %. Mean wind speeds, and percentage of calms (wind speeds < 0.5 m s−1 ) are denoted by m and c respectively. On the right, are corresponding diurnal temperature profiles. 61 Figure 3.3 Pollution roses as the relative contribution of wind sectors to mean peak season concentrations of PM2.5 , SO2 (summer), and NO2 (winter) at monitoring stations. Riverlodge’s PM 2.5 peak season is winter while it is in autumn for the other stations. Radials (%), are annotated on the wedges. Also shown are spatial distributions of hourly emissions 62 3.3.2 Model performance evaluations The adequacy of CMAQ to simulate observed air quality variables can be deduced in a relative sense, by comparing between the various PBL schemes, across monitor locations, for specified averaging periods and with model performance benchmarks (where applicable), using several statistical measures. Evaluations are conducted for each pollutant and are also analyzed graphically. SO2 SO2 simulation biases are according to location; the majority of PBL schemes overestimate Whitesail concentrations, and underestimate concentrations at Kitamaat Village and Terrace stations (Table 3.5). Across stations, NMB ranged -0.9 to +1.1, and MFB ranged -1.7 to +0.7. Out of 15 evaluations, FAC2 was deemed satisfactory in five, where all the PBL schemes are represented, except the UW scheme. NSD values closest to 1 at Whitesail suggest the best emulation of observed SO2 variability by CMAQ. Both Whitesail and Terrace are downwind of the major SO2 elevated source in the TKV during summer, but Whitesail is nearer. The diurnal peak around 1000 PST (Fig. 3.4a), is more delayed in modeled profiles for Terrace. Differences in emission mode, whereby fugitive, ground emissions are common in Terrace, likely accounts for discrepancies in the timing of peak concentrations. As the ground warms during the day and PBL deepens up to heights of buoyant SO2 plumes from the smelter source, more pollutants are brought into contact with surface air, leading to a quicker effect on simulated ambient levels for Whitesail. Indeed, PBL height (PBLH) maxima at Whitesail are more coincident with daytime peak 63 concentrations than at the Terrace station. Table 3.5 Model performance statistics for SO2 for various PBL parameterization schemes for the peak summer season. Values in italics indicate best performing locations per measure or schemes per location. Bolded values are within suggested benchmarks, only applicable to FAC2 and NMB. Measure Station MYJ MYNN3 UW ACM2 YSU Satisfactory (# out of 15) 0.5 ≤ FAC2 Kitamaat 11 19 0 4 3 ≤ 2.0 (# hours Whitesail 16 8 5 15 9 out of 24) Terrace 9 22 0 9 18 Kitamaat -0.15 -0.11 -0.25 -0.20 -0.20 Whitesail 0.21 0.65 -0.35 0.42 0.66 Terrace -0.41 -0.19 -0.64 -0.39 -0.33 Kitamaat -0.49 -0.34 -0.80 -0.66 -0.65 Whitesail 0.34 1.06 -0.58 0.69 1.07 Terrace -0.58 -0.27 -0.89 -0.54 -0.46 Kitamaat -0.65 -0.41 -1.34 -0.97 -0.95 Whitesail 0.29 0.69 -0.82 0.51 0.70 Terrace -0.81 -0.31 -1.62 -0.74 -0.60 Kitamaat 0.57 0.40 0.21 0.40 0.42 Whitesail 1.28 0.87 0.96 1.44 1.06 Terrace 0.53 0.63 0.27 0.60 0.32 MB (ppb) NMB MFB NSD 5 1 Daytime MYNN3, YSU and ACM2 outputs close to the stack emissions (i.e. at the White64 sail station) are mostly comparable. Probably due to generating lower wind speeds that are more in agreement with observations, concentrations maxima for these PBL schemes are greater than for the MYJ and UW schemes. YSU and ACM2 wind speed bias profiles are similar, which is likely because they are coupled to the same surface layer scheme (MM5-similarity). The other (TKE) PBL schemes are coupled to the Eta surface layer scheme, but it is evident that the general overestimation of onshore wind speeds is reduced in the MYNN3 scheme. Apart from having nonlocal convective mixing as with the YSU and ACM2 schemes, the MYNN3 scheme implements counter-gradient momentum diffusion (Olson et al. 2019) that mitigates output of excessive wind speeds. At the Terrace station, where strong winds are output from WRF, daytime MYNN3 wind speed bias approximates those of YSU and ACM2 schemes, thereby partly contributing to the MYNN3 scheme having the least SO2 concentration bias. Considerable nighttime overestimates with the MYNN3 and YSU schemes at Whitesail alludes to active, nonlocal transport in these schemes that are not inhibited by increased atmospheric stability. While the turbulence closure for YSU is purely nonlocal, the MYNN3 scheme includes a nonlocal turbulence production component for buoyancy effects of cloud-top cooling (Nakanishi and Niino 2009, Olson et al. 2019). The greater cloud cover amounts with MYNN3 (not shown), would enhance these effects. The strength of the vertical diffusion in a stable boundary layer is regulated by the buoyancy length scale and enhanced mixing with MYNN3 causes more downward transport of SO2 from aloft. Mixing-length formulations for both MYJ and UW schemes are strictly local, hence their limited capacity for pollutant transport from upper layers to the surface, despite slightly 65 Figure 3.4 (a.) Comparisons of modeled peak season (summer) SO2 profiles at Terrace and Whitesail stations with observations (OBS), and surface meteorological variables namely wind speed bias, air temperature bias, and PBL height (PBLH). Thin, black horizontal lines are zero bias axes. (b.) Vertical profiles of averaged potential temperatures at selected times, as atmospheric stability proxies for the various PBL schemes. The more the vertical homogeneity of potential temperatures, the less stable the atmosphere. Up to 2.8 ◦ C potential temperature change within the lowest 500 m is simulated at 0200 PST, Terrace station. more stable atmospheres than the YSU scheme (Fig. 3.4b). The ACM2 scheme does not produce as much nighttime concentration as the YSU scheme. For neutral, and stable conditions that dominate at night, the ACM2 scheme deactivates nonlocal transport of 66 scalar fluxes (Pleim 2007). Vertical exchanges via a purely local component thus result in pollutant concentrations comparable to the MYJ scheme. PM2.5 PM2.5 is underestimated (Table 3.6) in all simulations. Across stations, NMB ranged -0.8 to -0.5, and MFB ranged -1.4 to -0.5. Notably, modeled concentrations are not within a factor of 2 of observations (FAC2). In fact, none of the suggested performance benchmarks is achieved. For individual locations, the best performance is at Haul Road, apparently due to relatively low concentrations that reduce margins for bias. Ambient PM2.5 is least between 1000–1500 PST (Fig. 3.5). This should coincide with deeper PBL in this period but only the MYNN3 and MYJ schemes reasonably represent the inverse relationship between the daytime dip in station measurement and modeled PBL rise. Concentrations profiles for the other schemes are flatter, even when their PBLs rise less. PBLH variation influence on modeled concentrations, it seems, depends on the absolute height of the PBL itself, which also varies with season. At the Riverlodge station in Kitimat, where peak PM2.5 is during winter, PBLs are mainly thinner and less variable than the autumn Terrace profiles, also reflecting in weaker, diurnal variability of modeled concentrations. But it is also known that the pollution intensity at Riverlodge, where residential wood heating probably accounts for the wintertime PM2.5 peak, is less. The Kitimat area is more ventilated, with more evenly distributed diurnal PM2.5 levels. Less modeled concentrations at Riverlodge in comparison to Terrace, alludes to reasonable dynamical consistency of CMAQ simulations. 67 Table 3.6 Model performance statistics for PM2.5 for various PBL parameterization schemes in winter (at Riverlodge) and autumn (Haul Road/Terrace) peak seasons. Values in italics indicate best performing locations per measure or schemes per location. Measure Station MYJ MYNN3 UW ACM2 YSU Satisfactory (# out of 15) 0.5 ≤ FAC2 Haul Road 0 7 0 0 0 ≤ 2.0 (# hours Riverlodge 0 0 0 0 0 out of 24) Terrace 0 0 0 0 0 Haul Road -2.71 -2.35 -2.98 -2.73 -2.67 Riverlodge -3.52 -3.35 -3.94 -3.70 -3.42 Terrace -7.42 -7.15 -8.53 -8.06 -7.91 Haul Road -0.62 -0.53 -0.68 -0.62 -0.61 Riverlodge -0.65 -0.62 -0.73 -0.69 -0.63 Terrace -0.72 -0.69 -0.82 -0.78 -0.76 Haul Road -0.89 -0.73 -1.03 -0.90 -0.87 Riverlodge -0.97 -0.90 -1.15 -1.05 -0.93 Terrace -1.12 -1.05 -1.40 -1.27 -1.23 Haul Road 0.17 0.54 0.10 0.20 0.19 Riverlodge 0.20 0.31 0.06 0.09 0.24 Terrace 0.22 0.31 0.07 0.12 0.12 MB (µg m−3 ) NMB MFB NSD 0 0 0 Slight changes in the order of concentrations from the various PBL schemes, may owe to such spatial differences in-situ emissions. At Riverlodge, YSU simulates slightly higher 68 Figure 3.5 (a.) Comparisons of modeled peak season PM2.5 profiles at Terrace (in autumn) and Riverlodge (in winter) with observations (OBS), and surface meteorological variables namely wind speed bias, air temperature bias, and PBL height (PBLH). Thin, black horizontal line is zero bias axis. (b.) Vertical profiles of modeled potential temperatures at selected times, as atmospheric stability proxies with the various PBL schemes. The more the vertical homogeneity of potential temperatures, the less stable the atmosphere. Up to 3.0 ◦ C potential temperature change within the lowest 500 m is simulated at 0200 PST, Terrace station. concentrations than MYJ, but less at the Terrace station. Distinct from SO2 , PM2.5 emissions in the TKV are mostly from ground-level sources. Because YSU-predicted PBLH 69 at peak PM2.5 hours are roughly the same (about 350–400 m) at the two stations, smaller PM2.5 emissions around Riverlodge would entail weaker, vertical concentration gradient to the top of the PBL, therefore less nonlocal transport of pollutant away from the surface. Cooler temperatures (by ∼ 1 ° C) with the YSU scheme, possibly reducing the volatility of particulate constituents, may also have contributed to higher modeled PM2.5 than that of MYJ at Riverlodge. If nonlocal transport of PM 2.5 to higher layers is more effective at Terrace for YSU, then MYNN3 should produce less concentration at this station than the purely local MYJ scheme. Instead, the MYNN3 scheme not only outputs the highest PM2.5 but also better mimics diurnal levels than MYJ. With MYNN3, the evening peak is higher than the morning peak, in better agreement with the observation. A likely reason for higher output with MYNN3 than MYJ is the facility of counter-gradient fluxes that applies to any scalar quantity, in the MYNN3 PBL formulation (Olson et al. 2019). Further, the activation of CMAQ’s wind-blown dust module for all simulations may have contributed more fine particulates in model layers above the surface for counter-gradient transport with the MYNN3 scheme. NO2 Similar to PM2.5 , ground sources (motor vehicles, rail transportation, and other mobile emissions), account for NO2 at the Terrace station which is the sole, continuous monitoring location for ambient nitrogen oxides in the TKV. Model underestimation of wintertime levels (Table 3.7) is also evident. Across PBL schemes, NMB range -0.9 to -0.4 while MFB is 70 between -1.7 and -0.5. NSD values, as indicators of variability of modeled concentrations, are impressive for the MYJ and MYNN3 schemes, but MYJ has the best FAC2 rating. Table 3.7 Model performance statistics for peak winter NO2 with various PBL parameterization schemes at the Terrace station. Values in italics indicate best performing per measure. Bolded values are within suggested benchmarks, only applicable to FAC2 and NMB. Measure MYJ MYNN3 UW ACM2 YSU Satisfactory (# out of 5) 0.5 ≤ FAC2 ≤ 2.0 16 11 0 0 1 MB (ppb) -2.10 -2.64 -4.20 -3.54 -3.09 NMB -0.45 -0.57 -0.90 -0.76 -0.66 MFB -0.58 -0.79 -1.64 -1.23 -0.99 NSD 1.17 0.84 0.17 0.47 0.62 1 (# hours out of 24) 0 Diurnal pattern replication by the various PBL schemes (Fig. 3.6a) indicates reasonable success in matching emissions with anthropogenic origins. Simulations reproduce the morning and evening peak concentration periods, typical of urban vehicular traffic source of NO2 . Emissions underestimation may have caused the low biases in modeled concentrations particularly for overnight hours, but high hourly wind speed biases seem to contribute as well. Observed wind speeds at this station, situated in an area ringed by terraces, and where the cross-wise flow of the Skeena River channel relative to the TKV (see Fig. 1.1) complicates the atmospheric environment, averages 1.5 m s-1 in winter. Model overestimation of wind speeds by about 3.6 m s-1 on average, signifying greater than pre71 vailing outdoor ventilation, thus diminishes the skill of current PBL schemes to quantify pollutant concentrations around this deep-lying observatory. Figure 3.6 (a.) Comparisons of modeled peak season NO2 profiles profiles at Terrace with observations (OBS), and surface meteorological variables namely wind speed bias, air temperature bias, and PBL height (PBLH). Thin, black horizontal line is zero bias axis. (b.) Vertical profiles of modeled potential temperatures at selected times, as atmospheric stability proxies with the various PBL schemes. The more the vertical homogeneity of potential temperatures, the less stable the atmosphere. Up to 3.8 ◦ C potential temperature change within the lowest 500 m is simulated at 0200 PST, Terrace station. Better agreement of MYJ output with observation than the other PBL schemes differs from peak SO2 and PM2.5 concentrations for which MYNN3 values were greater. Atmo72 spheric residence times and abundance, chemical reactions, etc. vary between pollutants and can change between seasons. Daylight period is shortest in the winter season, and reduced solar heating of the ground affects the thermal structure of the boundary layer. The greater vertical variation in potential temperatures (Fig 3.6b) compared to other seasons (Figs. 3.4b, 3.5b), points to increased atmospheric stability for all PBL schemes, but the MYJ scheme is the most stable. This suggests that for modeled concentrations in the cold season, atmospheric stability is more important than differences in surface meteorological variables such as wind speed and air temperature amongst PBL schemes. An interesting aspect of NO2 levels from the various scheme relates to modeled PBLH. Figure 3.6a informs that the order of PBLH values (MYJ > MYNN3 > YSU > ACM2 > UW) is similar to that for NO2 concentrations. As stated in Chapter 2 (subsection 2.2.1), each turbulence closure scheme diagnoses PBLH based on certain criteria, hence their nominal values are essentially not comparable. Perhaps, their utilization within CMAQ could be for estimating vertical pollutant gradients between concentrations in the surface layer, and virtually zero concentrations in the free atmosphere above the PBL. The higher the PBLH, the weaker this gradient, resulting in lesser/slower pollutant transfer from the surface to higher layers. 3.3.3 Annual evaluations and spatial concentration differences Many air quality standards are based on top percentile concentrations of monitored pollutants. Therefore, the ability of model simulations to reproduce such levels at an annual scale is important to what confidence can be placed on their use for supplementing in-situ 73 monitoring. Figure 3.7 shows cumulative mean distribution plots of all hourly measurements and model outputs with the various PBL schemes. For SO2 (Fig. 3.7a) and NO2 (Fig. 3.7c), the mean of top quantiles of MYNN3, MYJ, YSU, and ACM2 PBL schemes are mainly the same orders of magnitude as observation data. These quite differ from PM2.5 where except for MYNN3, even the annual means of station measurements surpasses the concentrations for the top quantiles. This implies that PM2.5 model biases in other seasons hardly compensate autumn/winter underestimations. Modeled annual mean concentrations of air pollutants over the TKV domain is shown in Fig. 3.8. Spatial gradients are identifiable with pollutant source types, whereby the elliptical pattern for SO2 elevated point sources, differs from the more concentric contours for PM2.5 and NO2 . The discrete nature of SO2 emissions is evident in the maximum of average concentrations per PBL scheme (Table 3.8) that are 1-2 orders of magnitude above the annual station averages in Table 3.4. The coefficient of variation (CoV) per pollutant, however, does indicate that there is greater variability amongst PBL schemes with simulating these values for NO2 and SO2 than there is for PM2.5 . Smallest CoV with PM2.5 , thus, least concentrations sensitivity, suggests that the choice of a PBL scheme for an annual projection is more crucial for gaseous pollutants than for PM2.5 . The farthest travel of Kitimat stack emissions is by the MYNN3 scheme. With this PBL scheme, 0.5-1 ppb SO2 contours (Fig. 3.8) extend to Terrace, compared to about threequarters the distance with the fully nonlocal YSU PBL scheme. Innate approaches to nonlocal turbulent mixing in these schemes will account for these differences. The YSU scheme relies on entrainment at the top of the PBL (Hong et al. 2006). As pollutant concen74 Figure 3.7 Annual cumulative mean distributions of modeled, and actual pollutant concentrations at validation stations. The concentration at each octile/sextile is the mean of all hourly concentration values for it. Quantiles are according to the number of unique values in station data. Dashed vertical lines are the mean concentrations of station measurements in 2017. 75 Table 3.8 Highest simulated concentration of pollutants within the TKV modeling domain for various PBL schemes. CoV per pollutant, is the ratio of the mean to the standard deviation of values for all the schemes. Pollutant MYNN3 MYJ YSU ACM2 UW CoV PM2.5 (µg m−3 ) 7.3 7.6 5.9 5.8 6.8 0.11 SO2 (ppb) 44.0 22.7 23.5 19.6 4.4 0.55 NO2 (ppb) 4.2 3.4 2.5 2.2 2.2 0.27 trations approach zero in upper atmospheric layers, this process would less contibute to pollutant fluxes at lower layers, compared to the counter-gradient process in the MYNN3 scheme. Inland penetration of SO2 for the hybrid ACM2 scheme is less than the MYNN3 and YSU schemes. Indeed, that ACM2 yields an SO2 pattern that is comparable to the fully local, MYJ scheme, suggests dominance of stability conditions more permitting of layer-by-layer (local) transport. Pollutant levels are least with the UW, consequently the least spatial variability of modeled concentrations. Especially for SO2 , aerial gradient barely exists; the concentration hotspot is contiguous with the Kitimat industrial source. Because of relatively low boundarylayer heights modeled by the UW scheme, coupled with hot, forceful SO2 ejection from the smelter source, much of it probably is represented as directly transferred to model layers above the PBL, that is, to the free atmosphere. The UW scheme was designed for marine stratocumulus-capped boundary layers (Bretherton and Park 2009) that often are decoupled, consequently its consistent simulation of shallowest PBLs. Passive groundlevel PM2.5 and NO2 discharges within the PBL would also be subject to a strong vertical 76 Figure 3.8 Spatial distribution of mean annual concentrations of PM2.5 , SO2 and NO2 with various PBL schemes. 77 concentration gradient that enables rapid losses from the surface to layers aloft. Rather than mix, the UW scheme diffuse chemical species away from the surface layer (Bretherton and Park 2009), hence small outputs for all three air pollutants. 3.4 Discussion and conclusion The respective NMB and MFB ranges for peak season air pollutant modeling with different PBL schemes were -0.9 to -0.4 and -1.7 to -0.5 for NO2 ; -0.9 to -1.1 and -1.7 to +0.7 for SO2 ; -0.8 to -0.5 and -0.7 to -1.4 for PM2.5 . The lesser bias values per pollutant are close to, or within the range, that is reported in literature for CMAQ modeling over long durations (Hu et al. 2016, Kota et al. 2018, Syrakov et al. 2016). It needs mentioning, however, that much of recent CMAQ model demonstrations, including the abovementioned studies, are at ’super-regional’ scales (>1000 km) and similarities or differences in performance are not generalizable. Further, recommended benchmarks for regional atmospheric chemistry model performance (Boylan and Russell 2006, Emery et al. 2017) derive mainly from studies over US areas with higher background pollutant levels and more local sources. Apart from differences in grid resolutions and geography, there are differences in meteorological forcing data, choice of chemical mechanisms, number of validation stations, averaging periods and even model versions. Simulations in this study were performed at high resolution (1 km grid spacing) and used updated meteorologychemistry modules, hence reflect the current status of the WRF-CMAQ modeling system to emulate key air pollutants over a sparsely populated fjord valley. 78 While compensation of model biases across seasons may have led to better capture of top percentile concentrations for SO2 than for PM2.5 , peak season performances of individual pollutants highlighted temporal limitations of present-day PBL parameterizations. PBL schemes are usually developed and verified under warm, dry atmospheric conditions (e.g. Hong et al. 2006, Nakanishi and Niino 2009). The summer peak for SO2 coincides with a period of enhanced mesoscale events such as land-sea breezes, slope-valley winds, thus are better linked to in-situ contaminant distribution. The diurnal peak occurring at mid-day also suggested the important role of fumigation, caused by convective downward mixing of SO2 emitted from elevated stacks, on surface concentrations. This occurrence was reasonably mimicked by the various PBL schemes but with some differences that pointed to slightly better representation of pollutant mixing by some schemes than others. In colder seasons (autumn and winter) large scale circulations decoupled from the local terrain dominates, and simulations within a FAC2 for PM 2.5 and NO2 were fewer than for summertime SO2 evaluations. The MY-schemes outputs were frequently closest to in-situ pollutant measurements. For PM2.5 , MYNN3 was closest. Model output with the MYJ scheme was the best fit for the single-station evaluation for NO2 . PM2.5 and NO2 are predominantly emitted from ground-level sources. Both PBL schemes also performed well for SO2 , although on the evidence of good accuracy of maximum and minimum diurnal concentrations in the near field of stack emissions in Kitimat, the ACM2 and MYJ schemes better emulated temporal variations. Specifically, the MYJ scheme did not produce as much daytime fumigation as ACM2, nor excessive nighttime mixing and downward transport as MYNN3 and YSU. 79 However, being that MYJ implements only local mixing between model layers, less pollutant quantity at higher elevations, in comparison to YSU and MYNN3 schemes, was available for transport within the valley atmosphere. Therefore, for elevated pollutant point sources that may arise in the Kitimat area, nonlocal PBL parameterization schemes could be useful for projecting the most spread of emissions. The physiography of the study area may explain why outputs from the MY-schemes more closely matched with concentrations of ground-level pollutants. Due to proximity to, and regular inland advection of the marine boundary layer, Kitimat in particular, is subject to a shallow PBL. Further, there are many water bodies in the study domain, and much of the land surface is snow-covered for at least a quarter of a year. Moist surface and nearsurface conditions, probably restrain the formation of strongly convective regimes, and neutral and stable atmospheric conditions that are likely common in the valley are better represented by local transport formulations in the MY-schemes. These schemes indicated stronger vertical potential gradients than the other schemes, hence greater inversion tendencies that may have contributed to them producing the highest pollutant concentrations. The lower ranking of ACM2 performance was quite unexpected since it implements both local and non-local turbulence closure, depending on stability conditions and is the default scheme in the current version of CMAQ. Some limited-duration dispersion studies at coastal locations (Rakesh et al. 2013, Srinivas et al. 2015) have reported the suitability of MY-schemes. The MYNN3 and MYJ model runs however, required 1.4 and 1.1 times more storage space, respectively than YSU/ACM2. Thus, it seemed their deployment came with additional computing costs. 80 Modeled daytime concentrations that were within FAC2 at the major pollutant source areas—Terrace (NO2 ) and Kitimat (SO2 )—indicated CMAQ’s ability to provide reliable ambient estimates in the absence of fixed monitors. Better quantitative accuracy of the MY-schemes than others notwithstanding, the achievement of performance benchmarks for PM2.5 modeling was poor overall. Insufficient characterization of particulate emissions could have contributed to larger biases than for NO2 /SO2 . Specifically, it is likely that particulate emissions from residential wood stoves—a major source of ambient PM2.5 in the cold season in the interior of British Columbia—is underestimated in the emissions inventory. PM2.5 underestimations in chemical transport models are widely reported (Li et al. 2016, Hu et al. 2016, Tessum et al. 2015) but the Canadian context is against a backdrop of province-level compilation of area emissions. Lumping of emissions over large areas does not allow much flexibility with the SMOKE tool for simulations over domain sizes as small as the TKV. Air pollutant emissions disaggregated into provincial sub-units (e.g. regional districts), would enhance the spatial resolution of emissions input to the CMAQ model. Such emphasis may also benefit emissions inventorying via inverse modeling. While emissions are a source of modeling uncertainty, other factors could have undermined agreement between station records and modeled concentrations. Observations themselves are never perfect, and monitoring errors may arise with the aging of sensors, or for measurements at low ambient pollutant concentrations such as in the TKV. Observation error in pollution monitoring is acknowledged in several studies (Thunis et al. 2012, Pernigotti et al. 2013) and measurement uncertainty ranges are prescribed for differ- 81 ent pollutants (Gerboles and Reuter 2010). Representativeness error (Chang and Hanna 2004, Swall and Foley 2009) is also implicit when comparing point measures to grid cell volume-averaged estimates and cannot be eliminated. Further, random perturbations in the valley atmosphere contribute to modeling error. Smoothed topography in the meteorological simulations is also a plausible cause for modeled concentrations deviations from observations. This especially would be the case for PM2.5 modeling over Terrace which is in an area surrounded by mountains, and where the across-valley flow of the Skeena River complicates the wind environment. It is noteworthy that of the three air pollutants that were modeled, the top concentrations across simulations were most comparable for PM2.5 , hence the least sensitive to PBL schemes. It is important that updates to PBL parameterization schemes take cognizance of the potential for inherent biases associated with specific air pollutants, particularly over deep valleys to improve CMAQ modeling. 82 Chapter Four Gridded bias correction of modeled PM2.5 for exposure assessment and estimation of background concentrations over a coastal valley in northwestern British Columbia, Canada (In Press with the same same title at J. Air Waste Manage. Assoc.) Abstract Chemical transport models (CTM) can have large biases and errors when simulating pollutant concentrations. To improve the characterization of fine particulate matter (PM 2.5 ) over complex terrain for exposure assessments, three mathematical formulae that utilized the relationship between modeled and observed quantile concentrations at a monitor sta83 tion were developed. These were then applied to one year of CMAQ model output of PM2.5 over the Terrace-Kitimat Valley (TKV). The final products enhanced the representation of ambient levels at existing monitoring stations when evaluated with conventional statistical measures. Better agreement of corrected outputs with observed compliance metrics was also found. On average, the amended outputs had absolute errors of 11 % and 10 % for the annual mean PM2.5 and 98th percentiles of daily concentrations, respectively, compared to 45 % and 61 %, respectively, in the original output. These improvements provided greater confidence to use the reduced-bias outputs to estimate concentrations at locations without monitors. The fact that pristine conditions dominate the modeling domain was exploited to derive annual background PM2.5 concentrations of 2.0–2.3 µg m-3 over the valley. To the author’s knowledge, this is the first study to calculate background PM2.5 concentrations over northern BC coastlands through bias-correction of outputs from an air quality model. 4.1 Introduction Deteroriation of air quality due to particulate matter with an aerodynamic diameter ≤ 2.5 microns (PM2.5 ) is widely acknowledged. Exposure to PM 2.5 which may be released from natural and anthropogenic activities such as wildfires and residential wood heating, has been associated with adverse health effects, notably heart and lung impairment (Atkinson et al. 2015, Crouse et al. 2012). WHO (2016) estimates that globally in 2012, outdoor (ambient) PM2.5 was responsible for ∼ 3 million deaths in urban and rural areas. Because of morbidity and mortality impacts, regulations by various jurisdictions to limit ambient 84 levels are stringent, with ambient monitoring focused mainly on cities. However, investigations (Crouse et al. 2012) have also associated early deaths to long-term exposure to PM2.5 concentrations lower than previously thought to be harmful. Consequently, there is growing interest in qualifying pollutant exposure in the backcountry where there often is little monitoring, and underlying problems could remain undetected for long periods. The British Columbia Ministry of Environment and Climate Strategy has the mandate of regulating air emissions throughout the province, including developing standards and objectives on air quality that is protective of public health. They also partner with local communities in recommending planning goals for airsheds, particularly in defining desirable, voluntary limit values for ambient PM2.5 . However, goal-setting and overall air quality management in the TKV present some challenges. First, fixed-site pollutant monitoring stations are few, and much of the valley is not captured. Second, despite their regulatory intent, dispersion modeling realism in terms of identifying baseline airshed status is unknown. Past air quality modeling in the area (e.g ESSA Technologies et al. 2015) focused on predictions of contaminant concentrations arising from emissions changes. However, without establishing baseline pollutant concentration, it will be difficult to account for cumulative emission impacts and tracking progress in airshed management programs. Third, dispersion modeling may underrate or exaggerate the concentration of particulate pollution over an area. Unless bias correction (Porter et al. 2015, Neal et al. 2014) is performed, use of raw model outputs can misclassify ambient PM2.5 exposures. Finally, there is need to discriminate particulate levels that can be reduced, from those that cannot be mitigated. The latter, more appropriately referred to as background con- 85 centrations (McKendry 2006, Veira et al. 2013) is a vital element for consideration when setting air quality goals and mobilizing resources to achieve them. In this chapter, objective bias-correction formulas for CMAQ-modeled PM2.5 over the TKV, for the purpose of deriving baseline ambient exposure and background concentrations are developed and implemented. The subject of generating realistic concentrations of this pollutant at unmonitored locations, and in essence, estimating compliance to air quality standards are addressed. The remainder of this chapter is organized as follows. In section 4.2, data types, retrieval, and utilization for the development of corrective schemes, including the analytical methods are described. In section 4.3, the bias correction results are compared with the raw model outputs and with observations, and assessed for their closeness to indices of local regulatory compliance. Background concentrations PM2.5 are also derived. Section 4.4 discusses the relevance of final gridded concentrations for present and future airshed management and concludes the chapter. 4.2 Methods 4.2.1 Model simulations and observations data Demonstrations use PM2.5 outputs from the WRF-SMOKE-CMAQ modeling by the MYNN3 PBL scheme in chapter 3, which had the closest match with observations. The observations are valid hourly records from three PM2.5 monitoring stations (Table 4.1) for the same year of simulations and are available at https://envistaweb.env.gov.bc.ca/DynamicTable2.aspx? G_ID=327. Observational data completeness across stations were mostly > 90 % at sea- 86 sonal intervals. For the purpose of this study, seasons are defined as March-May (spring), June-August (summer), September-November (autumn), and December-February (winter). Table 4.1 PM2.5 monitoring locations, and hourly observation data completeness for year 2017, to the nearest percentage (%) Station Latitude ◦ N Longitude ◦ W Spring Terrace 54.522 128.608 100 96 98 62 89 Riverlodge 54.054 128.671 100 99 99 100 100 Haul road 54.029 128.702 97 98 98 99 98 Summer Autumn Winter Total Prior evaluations indicated significant quantitative differences of modeled PM2.5 levels from observations. Specifically, quantile distributions showed that the model underestimates the mean of the top quantiles by factors > 2. The errors in modeling that result in the need for bias correction are from two main sources. The first comes from inexactness of the emission inventory (amount, location, characteristics and configuration of emission sources); this directly impacts the modeled ambient levels. The other is associated with how well the model itself emulates atmospheric processes, which includes the representation of meteorological variables, and transport, dispersion, and transformation of pollutants. Consequently, an adjustment of concentration distribution in the raw output is necessary to ameliorate the above shortcomings. The original model representation of observations (see Fig. 4.1) is the basis for bias correction that is explained next. 87 4.2.2 Bias correction formulations Three separate mathematical equations relating observations to simulations are derived and analyzed for their effect on raw output: (1) simple linear regression; (2) power transformation (Li 2005) of the dependent variable in (1) based on a maximum likelihood parameter; and (3) second-degree polynomial. Figure 4.1 describes the three correction models. The equations are derived from data for the Terrace station. This choice for training the mathematical correction models is because the Terrace station reports the highest ambient PM2.5 , is affected by diverse pollutant sources (road dust, vehicular traffic, residential sources, etc.); and it is located in an area of significant terrain complexity (deeplying urban observatory). Half the dataset at this station is used to create the equations using values from every second day, and the remaining data are used for evaluations. Each of the equations is applied to the original hourly output for the entire grid to generate domain-wide corrections. In the rest of this paper, ’ORIGINAL’ refers to the original model output, while ’SIMPLE’, ’POWER’ and ’DEG_2’ refers to correction models (1), (2) and (3) respectively. 4.2.3 Evaluation measures Common statistical measures namely normalized mean bias (NMB), normalized mean square error (NMSE), and proportion of model results within a factor of two of observations (FAC2), are used to evaluate whether, and to what extent the final products reduce original biases. Several authors (e.g. Chang and Hanna 2004, Solazzo and Galmarini 2016) recommend these statistics for model performance evaluations and their formulae 88 Figure 4.1 (a.) Regression plots (quantile mean concentrations of model output against those of observational data) at Terrace, and derivation of bias correction formulae. Small open circles are the data points for both. (b.) Quantile plots for modeled and observed (OBS) data at pollutant monitoring stations. The concentration at each octile is the mean of all hourly concentration values within that octile. SIMPLE, POWER and DEG_2 are post-correction profiles. are in Appendix B. NMB and NMSE are bounded between zero and infinity, and values closer to zero are desired. Benchmarks for good model performance such as FAC2 ≥ 50 %, are also calculated. The assessment of modeled-observation data pairs is done for daily averaged data per season, following recommendations by Emery et al. (2017) for modeling of PM2.5 . Bias-corrected products are also assessed for fitness with prevailing air quality standards. Statistics for the latter are summarized as mean errors. 89 4.3 Results and analyses 4.3.1 Accuracy of bias corrected outputs Figure 4.1b showed the quantile mean concentrations of bias correction vis-a-vis the original simulation and observations across stations. The main effect of the regression schemes is to shift the distribution profiles closer to observation. Biases for top concentrations are reduced. However, mixed outcomes are evident; the mean of the top quantile remains underestimated at Haul Road but overestimated at the other two stations. The commonality of overestimation for the Riverlodge station and the Terrace station that is used for deriving the corrections, could be because both are near residential areas where PM2.5 emissions exhibit similar diurnal patterns. The Haul Road station in contrast, is more influenced by steady PM2.5 sources such as sea salt and industrial releases. The effect of bias correction at the seasonal scale is analyzed with paired data (Table 4.2). Whereas agreement with observation generally improves with corrections, that for the spring season is mostly more biased than the original output. This outcome nonetheless, is minor since springtime is not the peak season for ambient PM2.5 in the valley. Peak PM2.5 levels occur during autumn and winter seasons, and the corrections are quite effective for these periods. Specifically, post-correction NMB and NMSE for the Terrace and Riverlodge stations are smaller than with the raw output, although the negative biases for the autumn season remain. Importantly, all FAC2 and almost all NMB values for autumn and winter periods are within 50 % and ± 30 %, respectively, that are suggested by Emery et al. (2017) as criteria for good model performance. Notwithstanding, PM2.5 90 measurements vary from one location to another and the relevance of bias-correction in terms of matching compliance metrics is explored. Table 4.2 Statistical evaluation of bias-corrections in comparison to original model output, per season. Comparisons are based on daily means. Values in italics indicate where bias-corrected outputs are worse than the raw output. Spring Haul road Riverlodge Terrace Haul road Riverlodge Terrace Haul road Riverlodge Terrace Winter Terrace Autumn Riverlodge Summer Haul road Measure original -0.2 -0.2 -0.1 -0.5 -0.3 -0.2 -0.5 -0.5 -0.7 -0.4 -0.6 -0.6 SIMPLE 0.3 0.4 1.1 -0.3 0.2 0.8 -0.3 -0.1 -0.3 0.0 -0.4 -0.3 POWER 0.4 0.4 1.2 -0.2 0.2 0.8 -0.2 -0.1 -0.3 0.1 -0.3 -0.2 DEG_2 0.3 0.4 1.2 -0.2 0.2 0.8 -0.3 -0.1 -0.3 0.1 -0.1 -0.2 original 0.4 0.3 0.4 1.2 0.6 0.6 1.9 1.1 2.1 0.5 1.3 1.9 SIMPLE 0.8 0.7 1.4 0.6 0.4 0.9 1.4 0.6 0.6 0.7 0.8 1.2 POWER 0.8 0.8 1.8 0.5 0.5 1.0 1.2 0.6 0.6 0.6 0.6 1.2 DEG_2 0.8 0.8 2.0 0.5 0.5 1.0 1.3 0.7 0.6 0.6 0.7 1.3 original 76 78 70 60 74 80 54 58 24 72 32 39 SIMPLE 61 67 47 60 71 56 54 57 57 53 54 57 POWER 71 72 43 77 73 59 68 70 59 70 57 64 DEG_2 70 73 48 73 73 59 64 68 59 68 57 64 NMB NMSE FAC2 91 4.3.2 Evaluations for fitness with compliance metrics British Columbia has non-statutory objectives for ambient PM2.5 to protect human health and to guide decisions on the permitting of new or modified industrial emission sources (BCMOECS 2020). They prescribe a daily metric (DM) which is the 98th percentile of 24hour mean concentrations over one year (standard = 25 µg m-3 ); and an annual metric (AM) which is the annual mean of all hourly values over one year (standard = 8 µg m-3 ). These metrics are calculated for the original model concentrations and bias-corrected outputs and compared to corresponding calculations for observations, to assess fitness with prevailing regulatory objectives. Calculations (Table 4.3) indicate that bias correction improves the estimation of compliance metrics. Across the three stations, prediction accuracy of DM with bias-corrected outputs ranged between -10 % and +20 % (mean error = 10 %) as against -67 % and -62 % (mean error = 61 %) with raw output. Although air quality model evaluation literature does not suggest what qualifies for a good performance with percentile-based regulatory standards, a bias spread of ±10 % may be considered a desirable performance goal. Overall, the bias corrections with the different formulas are comparable. Notwithstanding, the higher-order regression models (POWER and DEG_2) have better comparisons for the Terrace observation data than SIMPLE, suggesting greater usefulness for fitting percentile-based metrics than a simple, linear scheme. For AM, bias-corrected values are within ±11 % of observations (mean error = 7 %) compared to between -42 and -50 % (mean error = 45 %) with the raw output. Because of the sign of model biases, the final gridded outputs would more likely overestimate DM than AM at unmonitored positions. 92 Table 4.3 Accuracy of corrected values for 24-hour (98th percentile of daily means) and annual metrics as normalized biases (in %). The mean error is the absolute ( | |) average of all post-correction biases Riverlodge Terrace Mean error Haul road Riverlodge Terrace Mean error Annual metric Haul road 24-hour metric original -62 -55 -67 -61 -42 -44 -50 -45 SIMPLE -10 12 12 11 -11 -5 7 8 POWER -6 19 1 9 -3 0 11 5 DEG_2 -7 18 3 9 -6 -9 9 8 Mean error 8 16 5 10 7 5 9 7 Also of interest is whether the corrections could be relied upon for indicating compliance with ambient PM2.5 standards. A decision-support framework outlined by the Canadian Council of Ministers of the Environment (CCME 2012) to categorize ambient PM2.5 exposures, and adapted to BC air quality objectives (ESSA Technologies et al. 2015), whereby a series of management actions are taken as air quality begins to deteriorate is presented for the TKV (Table 4.4). In applying this framework, exposure levels are color-coded according to concentation ranges of annual and daily metrics. Figure 4.2 describes the agreement among the original output, corrected values, and observational data. Different from what is obtained with the original values, all corrected values fall within the same categories as 93 observations. The strongest improvement appears to be for the Terrace station. Without reducing the biases, the raw outputs will indicate overly optimistic DM and AM compliances, which in turn, could undermine pollutant exposure. That corrections capture the same management category as observations, indicate reasonable ability to provide useful guidance at unmonitored locations. Table 4.4 Tiered air quality management threshold values and actions for ambient PM2.5 adapted to British Columbia from CCME (2012) Limits (µg m−3 ) Annual 98th percentile of mean 24-hour daily mean 8 6 4 Exposure Assessment Management objective High Achieve standard Moderate Prevent exceedance Mild Improve air quality Low Maintain good air quality 25 17.5 10 Figure 4.3 shows post-correction spatial plots of AM for all three schemes. Pollutant concentrations are quite similar among correction schemes, and are below 8 µg m-3 except for a small area (4–5 grid cells) just west of Terrace. Here, modeled concentrations attain maximum values of 18.3, 12.7 and 29.2 µg m-3 for SIMPLE, POWER and DEG_2 corrections, 94 Figure 4.2 Correspondence of original and corrected outputs with observed CCME 2012 PM2.5 management categories 95 respectively. Highest post-correction PM2.5 levels for the Terrace area are consistent with expectations for urban settings, but comparable levels are also around the Haul Road station, near the coastal outlet. This area is an industrial zone where particulate emissions from an existing aluminum smelter and allied services are probably a major contributor. Sideward spread of stack emissions is restricted by valley walls to the west of the industrial hub, hence a localized area that is closer to the limit values than the rest of the Kitimat area. Away from the Terrace and Kitimat areas, modeled AM is less than 3 µg m-3 . 4.3.3 Estimation of background concentrations Based on the close fit between bias-corrected output with observation for the annual air quality management metric, background PM2.5 concentrations in the valley can be estimated. Background concentrations are pollutant levels in the absence of local anthropogenic emissions, and represent the sum of concentrations arising from natural processes and those transported into an airshed from afar (McKendry 2006). Ideally, longterm pollutant monitoring at a pristine site would provide background concentrations, but in its absence in the valley, and based on local knowledge, an unmonitored, uninhabited position in the modeling grid, distant from main emission areas is used. This position (Fig. 4.4) that is currently devoid of anthropogenic PM2.5 sources and thus representative of background conditions, is about 30 km from Terrace and Kitimat. Figure 4.4 indicates that the mean annual background concentration of PM2.5 for the valley is 2.0–2.3 µg m-3 . The AM plots in Fig. 4.3 also hint at this range, but it is worth mentioning that pristine environments are not confined to within the valley alone. The 96 Figure 4.3 Spatial plot of bias-corrected mean annual PM2.5 concentrations for (a) SIMPLE (b) POWER and (c) DEG_2 mountain slopes and high-altitude areas are also wildernesses. On the western mountain ridges that are exposed to Pacific air masses, the minimum concentrations are 1.0, 1.8 and 1.4 µg m-3 for SIMPLE, POWER and DEG_2 schemes, respectively. Given the role of 97 transcontinental transport in redistributing atmospheric aerosol and relative isolation of the TKV area from significant anthropogenic North American sources, these minimums can be regarded as background concentrations affecting this part of the North American west coast. These concentrations would result from emissions over the ocean, including sea salt (although this source would have a larger impact area like the valley bottom that are within the MBL), contributions from Eurasian sources undergoing transpacific transport, and wind-blown dust within the domain. Episodic particulate transport due to large-scale, natural, seasonal emergencies might influence background concentrations. For example, the summer of 2017 was one of the worst forest fire seasons in the BC Interior in recent history, although coastal areas were mostly spared the heaviest smoke due to predominantly westerly atmospheric circulation. Emissions from wildfires are not included in the model emission inventory. Figure 4.4 Background PM2.5 estimation from bias-corrected annual mean centreline concentrations (see Fig. 2.1) 98 4.4 Discussion and conclusion The application of linear regression models to raw CMAQ model output improved the representation of ambient PM2.5 over the TKV. The final products from the mathematical formulations were generally comparable, with slightly better agreement of modeled and stations data for correction that is based on power transform of the original model output. Improvements are remarkable when it is considered that corrections were with data from an urban monitor. The POWER formula, for example, reduced autumn NMSE from 1.9 to 1.2 at Haul road, and from 1.1 to 0.6 at Riverlodge, representing 37 % and 45 % decreases, respectively, from the original modeled values. Djalalova et al. (2015) reported a 50–75 % reduction in absolute errors with a Kalman filter-analog combination over the contiguous USA. Nonetheless, it is difficult to compare the present study with error reduction demonstrations elsewhere because apart from dissimilarities in geographical settings, there are differences in statistical measures employed, including averaging periods, data collation and observation network densities. It is worth emphasizing that the majority focus of PM2.5 modeling bias correction literature has been on operational forecasting for much larger geographical areas, and with higher ambient concentrations than in the TKV. In China were ambient particulate measurements are much higher, Lyu et al. (2017) found for a suite of 12-km forecast correction techniques, reductions in normalized mean errors in the ranges of 7.4–19 % for 1-day lead time. Bias corrections in this study applied retrospectively to simulations on 1-km horizontal grids, thus demonstrates the efficacy of quantile-based statistical techniques to improve high-resolution chemical transport model outputs, over complex terrain with sparse monitoring. 99 Degraded springtime FAC2 and NMSE, alongside generally improved statistics for autumn and winter seasons following bias corrections, suggested error compensations that were beneficial to modeling accuracy for periods of peak ambient PM2.5 . Further, there was substantial improvement in the representation of compliance metrics for ambient levels. Since the time order of hourly concentrations was not critical to determining whether the TKV complies with existing PM2.5 standards, the approach of using regression models that were based on quantiles-mean is valuable to airshed management. Specifically, corrections reduced large biases for the percentile-based daily metric to an average error of 10 %. A comparable effect was also obtained for the annual metric. The significance of improved agreement of simulations with observations is more confidence in base PM2.5 contour maps that are generated. For epidemiological studies on community health effects of particulate matter, such spatial representations are indispensable, since fixed monitors have limited aerial coverage. Because instances of post-correction negative biases across stations was less for the daily metric (3) than the annual metric (5), projections with refined outputs may be more intense for acute (daily) exposure than for chronic (annual) exposure. This view especially follows from transforming ambient PM2.5 levels into an exposure risk ranking (Fig. 4.5) according to thresholds that are enunciated by CCME (2012) and applicable to British Columbia. Also suggestive is that a small area (∼5 km2 ) just west of Terrace may be in violation of the provincial PM2.5 standard. At this stage, it is not certain the number of persons that may be impacted if borne out in reality, since this zone is outside the urban core. Indeed, one shortcoming is that the assessment does not account 100 for effective exposure, which depends on the duration of outdoor activity by individuals. Nonetheless, the evaluation hints at the possibility of an on-going problem. Whereas collecting ambient data and public education on local air quality are appropriate for all areas, additional surveillance and analysis of trends can assist with confirming the actual source of the higher levels of PM2.5 . This may warrant ground-truthing over several months, to preclude costly regulatory measures that might not be required in the present. PM2.5 constituents are not summed in CMAQ boundary conditions file that dictate modeled levels. This necessitated the derivation of pollutant background from reduced-bias concentrations within the model domain. Estimated mean annual background PM2.5 concentration over the TKV was 2.0–2.3 µg m-3 which is comparable to the value of 2 µg m-3 suggested by McKendry (2006) for the entire BC province, and 2.5 µg m-3 reported by Vingarzan (2007) for all of Canada. This is also within the range of 1–4 µg m-3 estimated for the western United States (USEPA 1996) which like the TKV, is affected by trans-Pacific aerosol transport. It needs mentioning that these works pertain to large geographical areas encompassing diverse physical environments and eco-climatic zones. The lower limit of background concentration from the present study is 0.2 µg m-3 higher than the lower end of 1.8–2.5 µg m-3 range calculated for locations in the BC Interior (Veira et al. 2013). Although the coastal setting of the TKV is exposed to both maritime and continental air masses, the valley possibly does not have higher background levels than areas further inland. Perhaps, differences in estimates could be more due to the estimation methods adopted by the various authors. For example, Veira et al. (2013) was for locations in small cities and was based on analysis of wind sectors, for which trajectories were assumed 101 Figure 4.5 Color-coded classification of modeled PM2.5 ambient exposure risk from POWER scheme using thresholds in Table 5. (a) annual metric (b) daily metric. 102 to have passed over negligible upwind sources of anthropogenic emissions. PM2.5 background concentration in the present study is within the range of 1.7–3.8 µg m-3 by a study that was based on monitoring data of six remote rural locations in Alberta (Cheng et al. 2000). In fact, a recent dispersion modeling with the CALPUFF model in the TKV (ESSA Technologies et al. 2015) used background PM2.5 levels ranging 2.2–3.6 µg m-3 for postprocessing of output concentrations at individual receptor locations. These comparisons allude to the usefulness of bias-corrected outputs of atmospheric-chemistry model simulations over pristine regions to estimate background PM2.5 concentrations. The desired goal of air quality objectives in BC, especially within residential areas is that air emissions do not exceed 8 µg m-3 PM2.5 average concentration per annum. The mean annual background PM2.5 concentrations of 2.0–2.3 µg m-3 in this study, translates to 25– 30 % of the provincial limit. It is reassuring that 2017 ambient concentrations for a vast portion of the valley are close to this level, implying that a ’‘business as usual’ economic activity in the area should sustain good air quality. Because aerial exceedance of the provincial annual metric is also small, isolated issues can be dealt with as they arise. However, in the event that proposed industries become operational, or work camps are built in confined neighborhoods, the adoption of local target limits for ambient PM2.5 will be beneficial. Specifically, the Terrace area, where topographical and meteorological conditions seem less favorable to dispersion of pollutants, and where some of the additional industrial workforce in the TKV could reside, should be prioritized for particulate emissions abatement. Overall, it is important that airshed planning goals for the TKV, particularly for PM2.5 , recognize that although anthropogenic emissions are low for the 103 most part at present, the natural background constitutes a portion of the ambient level that is not amenable to control. 104 Chapter Five Acid wet-deposition modeling sensitivity to WRF-CMAQ planetary boundary layer schemes and exceedance of critical loads over a coastal mountain valley area of northwestern British Columbia, Canada (Also published at https://doi.org/10.1016/j.apr.2020.09.014 in Atmos. Pol. Res.) Abstract Computational tools used to implement the critical-load approach of atmospheric depo- 105 sition impact often do not elucidate modeling uncertainty, making it difficult for environmental policy-makers to know how much confidence to put in its results, also hampering aspects that may need improving. This study evaluated acid deposition modeling for various parameterizations of the planetary boundary-layer (PBL) over the Terrace Kitimat Valley (TKV). Of five schemes, simulations with the MYNN3 and MYJ PBL schemes best captured weekly wet deposition fluxes of acidifying ions (SO4 2− , NO3 − , NH4 + ) within a factor of 2 of observations at an industrial fence line station. Alongside the YSU PBL scheme, these two schemes slightly overestimated the chemical species at a station that is distant from major anthropogenic precursor sources in the valley, hence useful for worstscenario projections of atmospheric deposition on the environment. Forest soils in the vicinity of a large aluminum smelter in Kitimat was estimated to exceed the critical load of acidity by 30.1–53.5 kg S ha−1 yr−1 . Exceedance of critical load of nutrient nitrogen restricted to the Terrace area (≤ 7 km2 ) ranged between none and 0.71 kg ha−1 yr−1 . This work provides guidance for using PBL schemes in the Weather Research and Forecasting model that is coupled to a deposition model when calculating critical-load exceedance over temperate, rugged, coastal geographies. 5.1 Introduction Aside from degrading ambient air quality, ecosystem impairment is one major concern of the release of pollutant gases such as sulfur dioxide (SO2 ) and nitrogen dioxide (NO2 ). These gases, as well as ammonia (NH3 ), are precursors of acidic compounds which include sulfate (SO4 2− ) and nitrate (NO3 − ) that can be deposited in wet form or fall as 106 dry deposition, leading to the acidification of soils and surface waters in regions with shallow, base-poor soils and low mineral weathering rates (Driscoll et al. 2001, Sullivan 2000). Consequences of acidification include the accumulation of hydrogen ion (H+ ) and increase in aluminum (Al) concentration in soil water and surface waters, which can adversely impact forest ecosystems and lead to the loss of aquatic biota (Driscoll et al. 2001, Sullivan 2000). In addition to its acidifying role, increased levels of nitrogen (N) deposition can lead to eutrophication. Although considered to be a limiting nutrient for many ecosystems, large amounts of N may cause loss of sensitive floral species, alter forest productivity, and increase leaching of NO3 − to surface waters (Dise et al. 2011, Du et al. 2019). Adverse effects from acidification or eutrophication are usually not immediate due to the inherent buffering capacity of soils and time-lags in ecosystem response (Cosby et al. 1985). The critical-load concept is a useful approach that has been developed to assess whether pollutant deposition at any location is harmful. Critical load is a quantitative measure of acid- or eutrophication-buffering capacity of an ecosystem (Pardo 2010). It serves as an objective metric that can be used to determine both the spatial extent of a region being subjected to damaging levels of sulfur and nitrogen deposition, and the magnitude of the acidification (Moran et al. 2008, Pardo 2010). Acidification is likely to occur if annual total S and N deposition to an ecosystem exceeds the critical loads. Critical loads of acidity (CLA) and nutrient nitrogen (CLNnut ) is primarily controlled by soil bedrock and geology, and smaller values mean lower acid/eutrophication buffering capacity. CLA fields for forest ecosystems have been derived at national and regional scales and are being har- 107 nessed for priority protective actions. The combat of acidification problems in eastern Canada, for instance, has chiefly relied on such information in creating frameworks (e.g. CCME 2014a, ECCC 2018a) for the reduction of precursor SO2 and nitrogen oxides (NO x ) emissions. Meanwhile, plans for heavy industry in parts of western Canada have brought into spotlight the potential for ecosystem acidification and more nitrogen deposition. The use of computational models in quantifying acid deposition and critical loads, and calculating exceedances is broadly recognized as a useful management tool, in that it permits estimating total acid deposition at unmonitored locations. Such means was recently deployed in the TKV with the Lagrangian CALPUFF modeling system (ESSA Technologies et al. 2014, Williston et al. 2016), wherein the risk of direct and indirect impacts of SO2 and NOx on terrestrial and aquatic ecosystems for a range of future emissions scenarios was assessed based on CLA and CLNnut generated from steady-state models and empirical observations. Meanwhile, advanced 3-dimensional Eulerian platforms for simulating complex interactions between meteorology and atmospheric chemistry and pollutant transport and deposition, have been deployed elsewhere, however with little information on optimal physico-chemical options in such models. Specifically, simulations of acid deposition models could be sensitive to physics parameterizations such as planetary boundary-layer schemes in coupled meteorological models. In wet, humid climates and frequently foggy areas such as the TKV, the removal of pollutants from the atmosphere by water droplets could be non-negligible, and a model’s handling of moist turbulence is relevant to the process, hence the importance of examining the influence of PBL schemes on wet deposition. Several researchers (e.g. Appel et al. 2010, Queen and Zhang 2008, 108 Williams et al. 2017) have explored the validity of CMAQ model-predicted acidifying quantities, subject to constraints of aerial precision and temporal coverage. For instance, Queen and Zhang (2008) performed CMAQ simulations for wet acid deposition at 4-, 12- and 36-km horizontal grid resolutions over North Carolina, USA but for two different months. Indeed, there is a lack of research not only on sensitivity of acid deposition modeling to PBL parameterizations at a time scale of a year or more, but also in characterizing uncertainty of simulation ensembles at high horizontal grid resolution over complex orography. With increasing recourse to computational systems in environmental decision-making at the local scale, it is imperative that modeling options are evaluated at finer resolutions for long durations to ascertain their predictive capacities. In this chapter, wet deposition of acidifying species simulated by CMAQ version 5.2 for various PBL schemes in input meteorological fields are evaluated against observations for one year (2017) in the TKV. Indicators of how well wet acidic depositions at the stations are reproduced, are calculated for each scheme. Spatial differences in total acid depositions between schemes are examined. Portions of the valley that are likely impacted by deposition of acidifying emissions are also identified, and above critical loads are estimated. Hence, this work also aims at providing baseline acidification status upon which changes by future emissions can be tracked. The remainder of the chapter is organized as follows. In the next section, CMAQ deposition modeling set-up, including the PBL schemes that are tested, the model grids, chemical speciation and output post-processing are described. In section 5.3, results of simulations, alongside comparative evaluation of performance for the PBL schemes are presented. Critical load exceedance fields for S 109 and N deposition fluxes are also derived. Section 5.4 provides a discussion of the results, including spatial variations in contributions of wet deposition and dry deposition across modeling options. The final section (5.5) summarises and concludes the chapter. 5.2 Methods 5.2.1 Deposition simulations The numerical experiments to generate deposition fluxes with various PBL schemes namely the MYNN3, MYJ, UW, ACM2 and YSU PBL parameterizations is the same that was used for modeling major air contaminants in Chapters 3 and 4. Concisely, the simulation platform consists of three main components namely: (a) the Weather Research and Forecasting (WRF) model, version 4.0: a meteorological model, (b) the Sparse Matrix Operator Kernel Emissions (SMOKE), version 2.6: an emissions processing tool, (c) the Community Multiscale Air Quality (CMAQ) model, version 5.2: an atmospheric chemistry model that also calculates acid deposition. Hourly CMAQ simulations for atmospheric deposition were retrieved and post-processed with built-in utility packages and I/O API tools (Coats Jr. 2017) to obtain: accumulated wet deposition fields for evaluation against 2017 measurements in the valley (Fig. 5.1), and annual total-sulfur (total-S) deposition and annual total-nitrogen (total-N) deposi110 tion fields for CLA / CLNnut exceedance using critical loads reported in Williston et al. (2016). For comparison to weekly observed wet deposition data, precursor gas and particle phases of related output species were combined according to formulas provided for the CB05 chemical deposition mechanism. As an example, sulfate wet deposition included fine and coarse mode sulfate particles, and sulfur dioxide gas. To calculate annual acid deposition, wet and dry CMAQ deposition flux fields for S and N were summed over all hours in the year from precursor compounds. Although CMAQ outputs wet deposition separately from dry deposition, both computations involved similar set of gas-phase and particle species. Computation for total-N applied to a dozen chemical compounds including ammonium particles and ammonia gas, dinitrogen pentoxide, nitrous acid, nitrochloride and peroxyacetynitrates. Computation for total-S applied to fewer species. 5.2.2 Measurement data and performance measures Deposition records in the valley for 2017 were obtained from two National Trends Network (NTN) sites (http://nadp.slh.wisc.edu/data/sites/list/?net=NTN). The Haul Road station is in the industrial area of Kitimat, while the Lakelse Lake station is in the vicinty of a lake with the same name, 20 km south of Terrace. Both stations perform weekly sampling for acidifying ions (SO4 2− , NO3 − , NH4 + ), and also gauge precipitation volumes. Each measures deposition from precipitation through a continuously operating wet deposition collector. The collector opens when its sensor detects precipitation, allowing it to fall into a bucket, and closes when there is none. The collections are weekly after which 111 Figure 5.1 Left: nesting set-up for WRF-CMAQ deposition modeling over the TKV. Orange boundary specifies northwestern British Columbia region from which boundary concentrations to TKV’s domain (black) were derived, while the red boundary is TKV’s meteorological domain. The boundary condition domain is 60 × 40, at 5-km grid spacing. Legend is for elevation above sea-level in meters. Right: Enlarged TKV area identifying deposition monitoring stations (with markers). The Haul Road station in Kitimat is within 2 km of an aluminum smelter that emits large amounts of SO2 . The Lakelse Lake station is about 45 km away from the smelter and 20 km from Terrace 112 another clean bucket is provided. Each site is also equipped with a weighing bucket rain gauge to provide a continuous record of rainfall. Out of 52 weeks, measurements for chemical species were available for 43 and 41 weeks at Haul road and Lakelse Lake respectively. Precipitation chemistry and co-located precipitation amount data were used to calculate annual calendar year bulk SO4 2− , NO3 − , and NH4 + deposition rates in 2017: Deposition (kg ha−1 yr−1 ) = ∑(sub ppt * Cs /100), where sub ppt is weekly rain gauge reading (mm), Cs is concentration of SO4 2− , NO3 − or NH4 + (mg litre−1 ) in a sample. For evaluations, species concentrations at model grid cells corresponding to ground locations of observations were retrieved. In line with recommendations for use of multiple statistical indicators (Dennis et al. 2010, Solazzo and Galmarini 2016), the number of predictions within a factor of 2 of observations (FAC2), mean bias (MB), normalized mean bias (NMB), mean fractional bias (MFB), and Pearson’s correlation coefficient r were used to evaluate fitness between observed and modeled depositions. The formulas for MB, NMB and MFB are outlined in Appendix C. 113 5.3 Results and analyses 5.3.1 Performance evaluation for wet deposition of acidifying species Deposition measurements in comparison to simulations are indicated in Fig. 5.2. Modeled profiles have similar orders of magnitude as observations, and consistent with quantities of species (SO4 2− > > NO3 − > > NH4 + ). The deposition of SO4 2− is much underestimated at the Haul road station which may be due to several reasons. As previously indicated, this station is on the fence line of the smelter site SO2 source in Kitimat. The SO4 2− discrepancy possibly stems from discretization errors of I-km grid spacing near a major point emission, also considering the inverse relationship between ambient precursor concentrations and deposition. The total flux of pollutant discharge is a combination of ambient, deposited and transformed fluxes, and the more quantities are air-borne, the lesser would be the amounts that can be allocated to deposition by the model. Evaluations for air pollutants in the Kitimat area (Chapter 3), reported overestimation of SO2 , possibly resulting in less allotment to the deposition component. Substantial underestimation of the precipitation in the snowy months of November through February that contributes nearly half of total annual precipitation, occurs across simulations. Dry precipitation bias thus contributes to smaller wet SO4 2− deposition than is observed at this station. But even in instances where weekly precipitation is reasonably simulated by a PBL scheme, such as in March with the MYJ scheme, 0.3–1.8 kg ha−1 SO4 2− deposition underestimations occur, implying that estimates of other meteorological inputs to CMAQ are also crucial to the accuracy of wet deposition modeling. Modeled winds in particular, were 114 Figure 5.2 Time series of weekly observed wet deposition of NH4 + , NO3 − , SO4 2− , versus simulations from various PBL schemes at Haul Road and Lakelse Lake stations. Gray shading corresponds to the periods of heavy precipitation in the valley from November to March. Broken observation profiles are for weeks with missing measurement. stronger than field observations, with wind speed overestimates of 2.0–3.5 m s−1 in the cold, wet season. Because northerly winds are more frequent during winter, overesti- 115 mated wind speed would cause wet deposition of industrial SO2 emissions to be further coastward, away from the station, hence simulation biases that are larger than those of the warm, summer period. Winds are southerly in the summer months, but lesser precipitation that results in smaller wet deposition, in addition to more accurate precipitation and wind speed values, ensures better agreement of modeled SO4 2− deposition with the Haul Road station measurement than in winter. At the Lakelse station, 45 km from the smelter site and in pristine surroundings, modeled wet SO4 2− deposition is mostly overestimated even though precipitation is underestimated. Modeled SO4 2− deposition is positively biased with MYJ, MYNN3 and YSU schemes, and all PBL schemes have > 50 % of paired data within a factor of 2 of observations (Table 5.1). For NO3 − and NH4 + whose quantities are also smaller at the Lakelse station than at the Haul Road station, model errors, including for normalized biases are frequently less at the Lakelse station. For instance, the lowest NMB for NO3 − and NH4 + at Lakelse lake (Haul Road) are 0.01 (0.03) and 0.01 (-0.21) respectively. These comparisons suggest that the nearer the distance of a location to the precursor source, the more important the accuracy of meteorological inputs to CMAQ wet deposition modeling. The frequency of best statistical values should be useful in ranking of the performance of individual PBL schemes, especially for the pristine Lakelse location—a higher management priority from an ecological change perspective. For the three chemical species however, the best statistical scores are mixed and no one PBL scheme is consistent for all the measures. Further, a PBL scheme may demonstrate relatively good fitness for a specific statistic just because it is poor in another. The UW scheme for example, has the 116 best r scores for SO4 2− and NH4 + at this station due to negligible temporal variability of its small outputs. Given that acidic wet depositions distant from precursor emissions may vary from year to year, the inherent capacity of a PBL scheme to capture aggregate deposition, rather than time correlation is more pertinent to environmental protection. Overall, the MY-schemes (MYJ and MYNN3) output the greatestquantities of acidifying species in wet form, followed by the YSU PBL scheme. This outcome is related to how much precipitation is generated by the various simulations. Despite frequently producing the highest monthly precipitation and the biggest total precipitation, followed by the MYNN3 PBL scheme, it is mainly in September that the MYJ PBL scheme generates more chemical fluxes than the other schemes. This again raises the possibility of properties aside from precipitation amounts uniquely influencing atmospheric wet deposition in the various PBL schemes. Highest SO4 2− deposition with the MYNN3 PBL scheme is partly due to simulating the greatest amount of ambient water vapor (refer to Chapter 2) that augments wet deposition via pollutant washout by dew and fog. Unlike the MYJ scheme, the MYNN3 scheme includes a partial-condensation model for subgridscale clouds (Nakanishi and Niino 2009), consequently simulating cloudier atmosphere over the study area than with the MYJ scheme. The UW PBL is also designed for moist turbulent processes (Bretherton and Park 2009), and produces dense, low clouds; however, deposition is less favoured, mainly because of less precipitation amount, and greater chemical species loss from the lowest model layer with this scheme. 117 Table 5.1 Model performance statistics for NH4 + , NO3 − , and SO4 2− for various PBL parameterization schemes at Haul road and Lakelse Lake stations. Values in italics indicate the best performance at each location, per statistic. MYJ UW ACM2 YSU MYNN3 MYJ UW ACM2 YSU Lakelse Lake MYNN3 SO4 2− NO3 − NH4 + Precipitation Haul road FAC2 50.0 27.1 16.7 16.7 16.7 54.0 48.0 40.0 46.0 46.0 MB (mm) -8.6 -30.2 -33.7 -34.1 -33.8 -2.1 -11.9 -15.1 -13.8 -12.8 NMB -0.19 -0.66 -0.73 -0.74 -0.74 -0.07 -0.42 -0.53 -0.49 -0.45 MFB -0.15 -0.85 -1.00 -1.05 -1.05 -0.08 -0.53 -0.72 -0.64 -0.58 r 0.54 0.65 0.69 0.68 0.64 0.57 0.61 0.57 0.59 0.57 FAC2 57.5 50.0 25.0 35.0 35.0 43.9 46.3 43.9 53.7 51.2 MB (kg ha−1 ) 0.00 0.00 -0.01 -0.01 -0.01 0.00 0.00 0.00 0.00 0.00 NMB -0.21 -0.21 -0.57 -0.51 -0.43 0.13 0.10 -0.38 -0.18 0.01 MFB -0.28 -0.28 -0.86 -0.72 -0.58 0.05 -0.02 -0.57 -0.29 -0.07 r 0.29 0.28 0.15 0.18 0.26 0.33 0.38 0.50 0.43 0.34 FAC2 45.0 42.5 37.5 27.5 37.5 36.6 46.3 39.0 46.3 39.0 MB (kg ha−1 ) 0.01 0.00 -0.01 -0.01 -0.01 0.02 0.01 0.00 0.00 0.01 NMB 0.16 0.03 -0.26 -0.28 -0.17 0.59 0.42 0.01 0.08 0.33 MFB 0.09 -0.01 -0.38 -0.35 -0.21 0.39 0.24 -0.09 -0.01 0.19 r 0.12 0.15 0.10 0.08 0.14 0.23 0.32 0.40 0.43 0.31 FAC2 7.5 7.5 0.0 0.0 2.5 65.9 70.7 56.1 70.7 65.9 MB (kg ha−1 ) -0.59 -0.59 -0.66 -0.64 -0.63 0.01 0.03 -0.04 -0.01 0.02 NMB -0.81 -0.82 -0.92 -0.89 -0.87 0.15 0.29 -0.48 -0.12 0.19 MFB -1.35 -1.38 -1.72 -1.62 -1.55 0.11 0.16 -0.72 -0.21 0.12 r 0.11 0.33 0.29 0.21 0.24 0.66 0.66 0.85 0.78 0.73 118 5.3.2 Comparison of total nitrogen and sulfur deposition among PBL schemes Aggregate (wet plus dry) sulfur and nitrogen deposition were computed and Fig. 5.3 shows the spatial gradients for model runs with the various PBL schemes. Except for the UW PBL parameterization, outputs generally reflect anthropogenic sources of N and S deposition, namely urban and industrial activity, respectively. Modeled annual deposition is mainly < 5 kg ha−1 for S and < 2 kg ha−1 for N. Amounts tend to be greater on the ridges on the east side of the valley than on the west side, indicating reasonable accounting of the effect of regional circulations on deposition in the simulations. Eastmoving winds originating from the Pacific Ocean transport the valley’s atmospheric contaminants, some of which are eventually deposited on west-facing slopes. More affected by the west-east circulations are areas of steeper topography near Terrace (c.f Fig. 5.1), and around the coastal channel where the drift of marine clouds could be impeded by the fjord walls. But quite different from S deposition, elevated N deposition on the ridges around Terrace are comparable to that at its location in the valley. This suggests present N deposition due to anthropogenic activity is not much greater than occurs naturally. Figure 5.4 compares the output among PBL schemes, and makes clear that total N deposition for the year in the valley is in the order MYJ > MYNN3 > YSU > ACM2 > UW, while that for S deposition is YSU > MYNN3 > MYJ > ACM2 > UW. Differences in the locations of precursor emissions sources, hence seasonality of downwind impacts would account for varying ranking of N and S deposition intensities. At half-way the dis- 119 Figure 5.3 Spatial plots of annual total-N (top), and annual total-S (bottom) with the various PBL schemes. tance between industrial SO2 stack releases in Kitimat and Terrace, the YSU and MYNN3 schemes output 2–4 kg ha−1 more S deposition than those of the MYJ and ACM2 schemes. 120 The TKV interior is subject to S deposition when onshore winds prevail, which is mainly in the dry, summer season. Thus implicit is that annual S deposition within the valley is more affected by calculations for the dry deposition component, which in turn derives from ambient concentrations of precursor species. Comparatively higher summertime SO2 concentrations were simulated by the YSU and MYNN3 PBL schemes which could be due to nonlocal mixing, theoretically ideal for convective conditions (Cohen et al. 2015) being accounted for by these PBL schemes. But it may as well be due to lesser precipitation in the warm season that removes less SO2 from ambient air, thereby making it more available for transport by the YSU and MYNN3 PBL schemes. In colder seasons, at which time the MYJ scheme has greater advantage over others in emulating the wet conditions that occur, northerly winds prevail and advected air pollutants are more susceptible to wet removal. Consequently, the MYJ scheme yields the greatest N and S deposition around the coastal parts of the domain. Unlike wet deposition, dry deposition is continuous, hence dominates total fluxes in the vicinity of major emissions (Fig. 5.5). For S deposition largely traceable to smelter site emissions in the valley, the ratio of wet deposition to annual total deposition (rwet ) increases going from the south towards the northern part of the valley. The dry flux component of total deposition is proportional to ambient precursor concentrations and their attenuation with distance from the main source, increases the relative contribution of wet deposition with distance. Hence, except the UW PBL scheme that outputs little S and N, there is broad similarity across schemes for rwet values. In the central portions rwet for S deposition is 0.3–0.5, ∼ 0.6 over the Terrace area, and ≥ 0.8 over the surrounding moun- 121 Figure 5.4 Pairwise differences in annual total-N (top), and annual total-S (bottom). Each domain-wide plot comes from subtracting the output of one PBL scheme from another. For example, MYNN3 – MYJ means the output from MYJ subtracted from that of MYNN3, etc. tains, consistent with heavier precipitation at higher elevations. Moister conditions over large inland water bodies than surrounding land areas such as 122 Figure 5.5 Domain-wide contribution of wet deposition as a ratio of the total-N (top) and total-S (bottom) for individual PBL schemes. 123 at the Lakelse Lake station (Table 5.2) also results in rwet that is comparable to those of high elevation areas. For N deposition, precursor concentrations originating from diffused sources (such as roads, residences, natural processes) as opposed to steady, point sources for S, result in higher, but less orographically nuanced rwet patterns. The range of rwet for N deposition is 0.3–0.5 over Terrace, 0.6–0.8 in the central areas of the valley, and generally ≥ 0.8 on the mountains. Table 5.2 % contribution of wet deposition to S and N annual deposition for individual PBL schemes at selected locations in the TKV. S N Location MYJ MYNN3 UW ACM2 YSU Terrace 53.5 52.5 94.8 50.3 53.5 Lakelse Lake 81.3 73.7 86.0 70.7 73.5 Haul road 8.6 5.2 32.3 3.9 3.5 Terrace 30.7 27.0 71.3 29.8 34.6 Lakelse Lake 90.5 88.6 84.0 88.3 87.2 Haul road 74.9 62.1 87.9 61.3 61.9 5.3.3 2017 critical-load exceedances for forest ecosystems Having totaled S and N deposition fluxes by the various PBL schemes, exceedances of thresholds for acidification and eutrophication are assessed. The critical load information that is reported for the TKV area (Williston et al. 2016) is exploited. The average CLA as estimated from steady-state mass balance models in that study is 181 meq m−2 yr−1 124 which is equivalent to 29 kg S ha−1 yr−1 . The mass balance models used as critical limits, base cation to aluminum ratios (Bc:Al) of 1 and 6 for the predominant coniferous forests growing on mineral soils (65 % of the area), and deciduous forest tree species, respectively (ESSA Technologies et al. 2014, Williston et al. 2016). The same study deemed a critical load of 4 kg ha−1 yr−1 nutrient N (CLNnut ) as protective of semi-natural terrestrial habitats in the area, including lichen communities, based on a synthesis of published literature. Both the CLA and CLNnut are subtracted from modeled annual S and N deposition fields, respectively, thereby providing the spatial extent of critical load exceedances. Area-weighted exceedance magnitudes are also calculated. Figure 5.6 shows year 2017 gridded exceedance fields for the MYNN3 PBL scheme. Exceedance areas for CLA are clearly separate from CLNnut , and in both cases, confined to grids over, or in close proximity to major anthropogenic precursor emissions. CLN nut exceedance in the Terrace area is most likely from motor vehicle and rail switch yard nitrogen oxides emissions. CLA exceedance in the Kitimat area is exclusively from aluminum smelter SO2 releases. These representations indicate that acidifying/eutrophying depositions are for the most part, well below critical loads for the forest ecosystem. Similarity in spatial exceedance patterns (not shown) also exists for other PBL schemes except that the UW scheme produced no exceedance, as summarized in Table 5.3. Aerial CLA exceedances are greater than those for CLNnut ; the greater CLA exceedance mainly due to continuous, large point emissions, rather than diffuse urban emissions. Apart from the MYNN3 scheme, the only other PBL scheme that demonstrates CLNnut exeedance, is the MYJ scheme. For acidity estimates, the MYNN3 scheme produces the most spatial 125 exceedance while the YSU scheme yields the highest area-weighted intensity. Figure 5.6 Domain-wide exceedances of critical loads of acidity (CLA)(left) and nutrient N CLNnut (right) with the the MYNN3 PBL scheme. 126 Table 5.3 Exceedances of critical loads of acidity (CLA) and nutrient nitrogen CLNnut with various PBL schemes in the TKV MYJ MYNN3 CLA exceedance area (km2 ) 8 10 0 7 7 Area-weighted CLA exceedance (kg S ha−1 yr−1 ) 37.5 43.4 0 30.1 53.5 CLnut exceedance area (km2 ) 5 7 0 0 0 Area-weighted CLNnut exceedance (kg N ha−1 yr−1 ) 0.71 0.69 0 0 0 5.4 UW ACM2 YSU Discussion Weekly biases for modeled wet deposition of NH4 + and NO3 − at deposition monitoring stations were within 0 ± 0.01 kg ha−1 . The near-zero biases were indicative of minimal background atmospheric pollution in the region. Simulations in relatively clean environments may exhibit near-zero biases for N-containing species with respect to other modeling configurations. Regional CMAQ modeling by Qiao et al. (2015) reported NH4 + mean biases of 0 kg ha−1 at 12-km, and -0.01 kg ha−1 at 36-km resolutions within a nature reserve in China. It should be mentioned that episode length in their study was not a year, rather a three-month period (June to August). Guo et al. (2018) found very small biases in wet deposition (-8.3 × 10−4 kg ha −1 for NH4 + , -3.79 × 10-4 kg ha−1 for NO3 − ) at its sole validation station in Louisiana, USA over a period of one month. From an accuracy view point, NH4 + and NO3 − ) quantitative uncertainties in the present study suggest that the effect of choosing a PBL scheme instead of another, is not trivial for CMAQ-coupled 127 modeling of wet-deposited N species over complex terrain. For SO4 2− that was the most prevalent of the three acidifying agents, normalized weekly biases ranged from a maximum underestimate of -0.92 at the fenceline of major industrial source of precursor SO2 emissions to a maximum overestimate of +0.29 at the pristine location further away. The majority of comparisons at both locations underestimated field measurements, for which underestimated precipitation contributed. But since some of the model runs overstated SO4 2− at the background (Lakelse) location, precipitation underestimation was possibly not the sole cause. Big penalties can arise for concentration fields with tight gradients that are slightly displaced from observations in time or space, and compensating errors at high (1 km) spatial resolution may have influenced deposition modeling performancee. Overestimation of SO2 in the near source area of stack emissions may have led to less quantity available for wet deposition. Uncertainty regarding the production and transformations of aerosol species could also have contributed to negative biases. Nearness to a maritime environment is a major influence on levels of atmospheric oxidants essential for conversion of SO2 to SO4 2− and CMAQ’s handling of chemical processes is the current science. Low bias for sulfate modeling is however, not unusual in industrial zones. Cho et al. (2017) noted 43 % underestimation in CMAQ simulated wet SO4 2− deposition at the Patricia McInnes site in the Alberta Oil Sands Region (AOSR) of western Canada at 4-km resolution. The majority emulation of sulfate deposition at Lakelse lake station in the present study, while mimicking the low levels of other acidifying agents, suggested that longrange deposition to pristine areas from elevated point sources of air pollutants can be reasonably predicted using 1 km horizontal 128 gridding. In terms of performance, the PBL schemes whose outputs closely matched acid wet deposition species, were the MYNN3, MYJ and YSU PBL schemes. These schemes tended toward moderate overestimation of chemical species in the far field of emission sources, hence could be valuable for worst-scenario projections of atmospheric deposition on pristine environments. It is recalled from investigations for major air pollutants (Chapter refchpt3 ) that the MYNN3 and MYJ PBL schemes yielded concentrations that were closest to peak ambient levels. Consequently, these two schemes have greater contaminant baselines for wet deposition than the other PBL schemes. Both the MYNN3 and MYJ schemes (MY-shemes) are turbulent kinetic energy (TKE)-based schemes, and the formulation of MYNN3 is similar to MYJ but with a higher level of closure (2 for MYNN3 compared to 1.5 for MYJ). The MYNN3 scheme’s influence on deposition in the TKV particularly derived from its consideration of moist convection, enabling additional pollutant settling via mists and cloudy boundary layers. The strength of the MYJ scheme, on the other hand, was due to its yield of the least precipitation deficit on an annual basis. Variants of the MY-schemes have been advanced for operational weather forecasting (Olson et al. 2019) and pollutant transport and dispersion mapping (Miao et al. 2007, Srinivas et al. 2015) in coastal topographies. The YSU scheme’s overestimations, especially for sulfate, owed to the position of the major pollutant source (aluminum smelter) that more opportuned long-range deposition by onshore winds in the warm season, when precipitation bias is also less.Unlike in the ACM2 scheme where nonlocal transport of meteorological quantities in the PBL is active 129 only during unstable conditions, nonlocal transport is the default in the YSU scheme. The wholly, nonlocal eddy diffusion approach of the YSU scheme thus was more supportive of prolonged mixing across model layers than the ACM2. By modeling greater vertical spread of emitted pollutants, chemical species are longer-lived and travel farther, hence more positive model biases at the pristine station with the YSU scheme than with the ACM2 scheme that is the default parameterization in CMAQ version 5.2. The UW scheme consistently produced the largest biases, apparently due to too little generation of precursor pollutants in the first place. Away from the Terrace and Kitimat urban areas, annual deposition fluxes were mostly comparable among the MYNN3, MYJ, YSU and ACM2 PBL schemes, also consistent with deposition fluxes in remote areas. The highest annual N deposition of 4.71 kg ha−1 (from the MYJ scheme) is in the lower end of the range for rural areas in Canada (4.3–11 kg N ha−1 , Zhang et al. 2009). That study was for sites in central and eastern Canada that are impacted by industrial NO x emissions from the US. In the present study, only the higher emission areas of Terrace and Kitimat evidenced predominant dry N and S deposition, linked to road and rail transport (NO x ) , and smelter (SO2 ) emissions, respectively. Wet deposition accounted for acidic deposition for much of the study domain. Excluding the UW scheme, percentage contribution of modeled dry deposition to total N at the Lakelse Lake station that is a pristine location, ranged 9.5–13 % as against 10–48 % for the rural Canadian sites study. However, should anthropogenic NOx , SO2 and NH3 emissions increase within the TKV, the contribution of dry deposition to total N and S depositions may also increase. Eularian grid modeling in areas of western Canada with large indus- 130 trial NOx and SO2 emissions, such as in the AOSR (Cho et al. 2017, Makar et al. 2018), have reported considerable aerial contribution of dry deposition to total deposition. Cho et al. (2017) estimates roughly 50 % of total S deposition to be due to dry deposition in the AOSR, and marginally greater than wet deposition in future projections. The wet deposition ratios and patterns in the present study indirectly inform the proportion of dry deposition which is a component that is difficult to measure in-situ and in real-time. Excluding results from the UW scheme, which likely is unrealistic on account of account of continuous smelting activity (> 60 years), area-averaged CLA exceedance was between 30.1–53.5 kg S ha−1 yr−1 over 7–10 km2 in the vicinity of the major SO 2 source in Kitimat. The CLNnut exceedance ranged 0–0.71 kg N ha−1 yr−1 over 0–7 km2 around Terrace. Spatial exceedances , including the largest (by the MYNN3 PBL scheme) were limited to the valley bottom. In contrast, CMAQ-derived critical load exceedances over the Georgia Basin of southern BC (Nasr et al., 2010), comprising areas of strong surface inhomogeneity as the TKV, and receiving modest precipitation (1200–1500 mm annually), were indicated to be greatest at high elevations and along steep valley slopes. Levels of socio-economic activity including much smaller industrial and urban S- and N-bearing emissions in the TKV than in the Lower Fraser Valley would suffice as to why upland areas are not impacted by the present emissions. Predictive modeling will be relevant to knowing whether anticipated emissions increase in the TKV and attendant deposition will enroach upland forests. 131 5.5 Concluding remarks CMAQ acidic wet deposition modeling sensitivity to parameterized planetary boundarylayer processes over the TKV found outputs to be comparable and credible for various selections except with the UW scheme that produced very low quantities. The validity of the majority simulations for wet fluxes, especially NH4 + and NO3 − was evident in their generation to same order of magnitude as observations, as well as being spatially correspondent with anthropogenic emissions and precipitation amounts. Overall, model fitness was better at the Lakelse monitoring station than at the Haul road station; slight overestimates at the former with the MYNN3, MYJ and YSU PBL suggesting reasonable fitness of these schemes for future projections. Domain-wide critical load exceedance calculations were then performed on modeled total depositions, one for annual sulfuronly deposition and another for annual nitrogen-only deposition. Consistent with local knowledge, small areas in the vicinity of major anthropogenic emissions of SO2 and NO x were estimated to be in exceedance of critical loads of acidity and nutrient N deposition. 132 Chapter Six Intercomparison of atmospheric datasets and PBL schemes for precipitation downscaling over a coastal mountain valley of northern British Columbia, Canada Abstract Modeling environmental futures of coastal areas demands processing and linking atmospheric data in a way that adequately represents the role of hydrology on ecological health. This study examined the capacity of dynamical downscaling to generate reliable precipitation fields over the Terrace–Kitimat Valley, an industrializing corridor in the Coast Mountains of northern British Columbia, Canada. Precipitation modeling un- 133 certainty was explored via year-long, 1–km resolution WRF model simulations for three atmospheric datasets and two planetary boundary layer (PBL) schemes. Underestimation by more than 40 % on average across stations with total precipitation ranging 1170– 2380 mm was found for simulations using either the North American Regional Reanalyses for atmospheric forcing or the Mellor-Yamada-Nakanishi-Niino level 3 (MYNN3) parameterization as PBL scheme. Persistently low bias from model configurations using these settings suggested that merely selecting an alternative atmospheric forcing dataset would hardly ameliorate systematic error occasioned by a poor choice of PBL parameterization. Hence, choice of PBL scheme and meteorological dataset are important when a numerical weather model is used for spatial estimates of precipitation. Model outputs best corresponded with annual gauge measurements when simulations with the MYJ PBL scheme were forced with ERA5, outperforming the North American Mesoscale Analyses (NAM_ANL). The latter however demonstrated better spatial and temporal fitness than ERA5. Using both datasets therefore may be valuable for projections related to environmental change. With either NAM_ANL or ERA5 for atmospheric forcing and MYJ as a PBL scheme, the uncertainty in annual simulated precipitation amount ranged between 40 % overestimation and 20 % underestimation of observational data. 6.1 Introduction The amount and distribution of precipitation are useful eco-climatic identifiers. Much of the world’s forests are in regions of moderate to high precipitation. A key component of the water cycle, precipitation is also a principal medium for the removal of pollutants 134 from the atmosphere to the surface. Despite the importance, in-situ precipitation measurement is hampered by expensive gauge installation and operating costs, and poor siting or relative inaccessibility of favored locations (e.g in remote areas). Additionally, spatiotemporal discontinuity of precipitation, often warrant reporting at intervals that may be overly aggregated to serve other purposes. As an example, without further processing, precipitation data archived as daily totals would be incompatible with numerical schemes requiring meteorological data at sub-daily/hourly time steps. Meteorological datasets with more spatial coverage and temporal resolution of precipitation values, provide a means to alleviate the challenges of poorly gauged catchments and aggregated reporting. While some datasets are interpolations of gauge measurements over grids, others employ sophisticated assimilation techniques that combine various observation sources (weather stations, buoys, aircrafts, radars, and satellite products) with numerical weather forecasts. Many of them can be accessed at no financial cost (https://rda.ucar.edu/). As meteorological datasets become more accessible, inquiries about their validity have also increased. A review of global precipitation datasets by Sun et al. (2018) found discrepancies of up to 300 mm yr−1 among products, with regional variability in the differences. Another global assessment with 22 gridded datasets by Beck et al. (2017) found gauge-corrected datasets incorporating daily data better calibrated a hydrological model than using temporally coarser gauge data. They also ranked the fitness of uncorrected datasets. Belo-Pereira et al. (2011) compared four global gridded datasets to observations within the Iberian Peninsula, and noted that some products better represented the annual cycle and drought than others. Significant differences among 135 precipitation datasets have also been reported for long term comparisons in Canada (Lilhare et al. 2019, Wong et al. 2017), emphasizing the importance of choice of gridded meteorological products for precipitation estimation. Selecting al dataset is more important when precipitation fields are required to address local environmental issues such as atmospheric deposition of pollutants. State-of-thescience meteorological drivers such as WRF can downscale global and regional datasets to much smaller domains. In addition to defining initial and boundary conditions, their implementation requires user-defined choices for nudging coefficients, physics parameterizations, nesting ratios, grid resolutions, etc. Several investigations of precipitation simulations in relation to these additional dependencies have been conducted. Li et al. (2014) assessed the role of physical parameterization scheme and model resolution on simulation of summer rainfall over the south-eastern US. They found convective rainfall to be more sensitive to planetary boundary layer (PBL) schemes than microphysics schemes, also inferring that the choice of cumulus scheme is more important than increase in model resolution. Hu et al. (2018) reached similar conclusions for the Great Plains region of the US where differences were more profound with change in cumulus parameterization schemes than other physics schemes. For snowfall over the northwestern Iberian Peninsula, Fernández-González et al. (2015) inferred better choices of microphysics and PBL schemes. Whereas physical parameterization and hydrological downscaling sensitivities are widely recognized as integral to reconstruction of catchment climatologies, much attention has been for short-term forecasts (Lee et al. 2015, Moya-Álvarez et al. 2020), extreme events (Pontoppidan et al. 2017, Remesan et al. 2014, Sarmadi et al. 2019) and 136 parametric uncertainty (Tian et al. 2017, Yang et al. 2012). Little work has explored precipitation sensitivity to atmospheric datasets and computational physics schemes from a common grid resolution. In this chapter, numerical simulations of precipitation for a full year over a coastal valley region of northern British Columbia (BC), Canada, are examined. Evaluations of outputs of dynamical downscaling with three atmospheric forcing datasets, and two PBL schemes in WRF are conducted with the aim of establishing the appropriateness of precipitation fields for potential deposition modeling from various combinations of forcing data and PBL parameterizations. This study therefore addresses the three following research questions: 1.) How well do individual simulations reproduce observed spatial and temporal distribution of precipitation in the area? (2) Which forcing dataset(s) and PBL scheme yield precipitation amounts most consistent with observations at individual stations? (3) What is the uncertainty in the precipitation field that comes with the best fit simulations? The answers to these questions are sought by evaluating model outputs with several performance measures and analytical plots. The remainder of this chapter is organized as follows. In section 6.2, precipitation measurement data, meteorological model configurations and driving atmospheric datasets, and the data analyses methods are described. Precipitation outputs from the simulations and comparisons to observations are presented in section 6.3. A discussion of dynamical downscaling sensitivities appears in section 6.4. The last section (6.5), concludes the study. 137 6.2 Methods 6.2.1 Description of datasets The selection of gridded products was based on their availability for the region at subdaily frequency. Sources providing data at longer time steps (e.g. monthly) or at very coarse spatial gridding (> 32 km) were not considered. Another consideration was retrieval in formats compatible with the WRF code. Lastly, the evaluations were for two major categories of assimilated products, namely analyses and reanalysis datasets (https:// rda.ucar.edu/datasets/ds083.2/docs/Analysis.pdf). For these reasons, the North American Mesoscale Forecast System’s Analyses (NAM_ANL: Environmental Modeling Center 2017), the North American Regional Reanalysis (NARR: Mesinger et al. 2006) and the fifth generation of the European Centre for Medium-Range Weather Forecasts (ECMWF) atmospheric reanalyses of the global climate (ERA5: ECMWF 2019) were selected. Procedures for creation of each dataset are not within the scope of this chapter; however detailed descriptions can be found in the cited literature and in references for the overview that is presented below. The National Centers for Environmental Prediction (NCEP) uses WRF to produce different mesoscale forecasts products and analyses over North America. The 6-hourly analysis dataset on 12-km horizontal grids (NAM_ANL) that is used however has a spatial coverage that excludes Canada north of 60 ◦ N. (https://www.emc.ncep.noaa.gov/mmb/ namgrids/g212.12km.jpg). Upgrades including transitioning from a 12-hour, to a 6-hour data assimilation cycle with hourly updates was recently implemented for this product 138 (Environmental Modeling Center 2017). The NARR is a 3-hourly product for North America developed by the NCEP. The NARR system uses the Eta 32-km atmospheric model with 45 vertical layers and a three-dimensional variational data assimilation approach. Precipitation is assimilated from regional hourly/6hourly multi-sensor (radar+gauges) precipitation analyses produced by 12 river forecast centers. (Lin and Mitchell 2005, Mesinger et al. 2006). While 1/8 ◦ daily rain gauge data analysis is used for the conterminous US, a 1◦ rain gauge analysis is used for Mexico and Canada (Shafran et al. 2004). ERA5 that has been used in previous chapters is a recent global reanalysis from ECMWF. It is produced from high-resolution model forecasts on a regular grid, and a ten-member ensemble, four dimensional variational data assimilation. Atmospheric data generated with 137 model levels are interpolated to 37 pressure levels up to 0.01 hPa. ERA5 is available at hourly intervals and a horizontal resolution of 0.281 ◦ (31 km) (ECMWF 2019). 6.2.2 Modeling configurations and experiments The WRF model (version 4.0) is used for simulating precipitation from the three atmospheric datasets for two PBL parameterizations. For turbulence parameterizations, the MYJ and MYNN3 PBL schemes were selected because they best quantified ambient concentrations of air pollutants in the valley (refer to Chapter 3). The WRF domains are set up with a nesting configuration in such a way that the innermost domains are identicallysized over the TKV (Table 6.1) and horizontal grid spacing is 1 km for all three datasets. To accomplish this, downscaling ratios of 1:5 for NARR and ERA5, and 1:3 for NAM_ANL 139 were used. Grid size of parent domains were 25 km for NARR/ERA5 and 9 km for NAM_ANL, hence intermediate domain sizes were 5 km and 3 km, respectively (Fig. 6.1). All domains are composed of 40 vertical pressure levels with the top level set at 50 hPa for NAM_ANL and ERA5, and 100 hPa for NARR. The same physics options that were used in Chapter 2 (section 2.2.2) were deployed for simulations, spanning 00:00 UTC 20 December 2016 and 00:00 UTC 2 January 2018. Similar to the previous procedure, each model run was initialized monthly, with model spinup of 1 day, and a further 1 day at the end of each month. The monthly overlap days, including for the period 20–31 December 2016 were discarded when merging hourly output files and only outputs for 2017 were retrieved. Table 6.1 Domain set-up and nesting with NAM_ANL (left) and ERA5/ NARR (right) Domains (dimensions) NARR / ERA5 NAM_ANL Parent (W-E x N-S) 100 x 100 100 x 100 Parent grid center 54.200 °N, 128.600 °W 54.200°N, 128.600°W 1st nested (W-E x N-S) 121 x 121 121 x 121 1st nested grid center 53.850 °N, 128.795°W 54.158°N, 128.671°W 2st nested (W-E x N-S) 101 x 121 100 x 121 2st nested grid center 54.223°N, 128.640°W 54.213°N, 128.695°W 140 Figure 6.1 Domain set-up and nesting with NAM_ANL (left) and ERA5/ NARR (right) 6.2.3 Observational data and evaluation protocol The WRF outputs are evaluated against station observations (Table 6.2) for 2017. Precipitation monitoring at the Lakelse station is as a part of atmospheric deposition sampling and uses the OTT Pluvio2 gauge with wind shielding. The other stations use Meteorological Service of Canada (MSC) Type-B rain gauges that are unshielded, with a collection rim height of 40 cm above the ground, and typical measurement uncertainty of ± 0.2 mm per measurement. Conventional statistical measures including for biases and errors, and Spearman’s rank correlation coefficients for monthly accumulations (Table 6.3) are used to assess fitness between model outputs and observations. Spatially-normalized plots are also used for domain-wide model outputs, and station-paired observances of precipitation amounts. Analyses are also performed with predictive scoring schemes. These are equitable threats 141 Table 6.2 Observational data locations for evaluation of precipitation simulations. Gauge stations include those operated by Environment and Climate Change Canada (ECCC) and BC Ministry of Forests, Lands, Natural Resource Operations, and Rural Development (BCMFLNRORD) Station Latitude Longitude Elevation Operating Collection (◦ N) (◦ W) (m) Agency frequency Terrace 54.501 128.625 58.2 ECCC Daily Lakelse 54.377 128.576 111.0 Rio Tinto Weekly Riverpark 54.170 128.577 75.0 BCMFLNRORD Hourly Kitimat 54.054 128.634 128 ECCC Daily score (ETS), probability of detection (POD), the false-alarm ratio (FAR), and the frequency bias index (FBI). POD is the rate of correct forecast of precipitation events (range 0–1 and a perfect score of 1). FAR is the rate of false positives (range 0–1 and a perfect score of 0). ETS is the rate of correct event forecasts, adjusted for correct detections that would be expected because of random chance (range -1/3 to 1, with perfect, and no skill scores of 1 and 0 respectively). The FBI gives the ratio of the estimated to observed precipitation frequency (range 0 to ∞ and a perfect score of 1). These detection metrics are often used in precipitation verification literature (e.g Peña-Arancibia et al. 2013, Schirmer and Jamieson 2015). Appendix H explains how these metrics are calculated. 142 Table 6.3 Fitness measures based on annual precipitation values for model outputs (M) and observations (O). In the formulae, rmi (roi ) are ranks of modeled (observed) amount for each month, hence n = 12 Statistical measure Formula Perfect value Bias (B) M-O 0 mm Error (E) |Bias| 0 mm Percent bias (PBIAS) M-O/O × 100 % 0 Percent error (PE) |M-O/O| × 100 % 0 n 6 Spearman’s correlation coefficient rk 6.3 1− ∑ ( r mi − r oi ) 2 i =1 n ( n2 −1) 1 Results and analyses 6.3.1 Quantitative biases for datasets and PBL schemes Recorded precipitation totals in the valley in 2017 varied from 1168 mm at the Terrace station in the north, to 2381 mm at the Kitimat station in the south; peak snow contributions (Fig. 6.2) of 25–40 % occured in March and November. Roughly one-third of the total precipitation amounts are during the months of October and November. Model outputs for total precipitation in 2017 at station locations, indicate PBIAS that are within ± 60 % of gauge measurements for all simulations (Table 6.4). The mean of PE (MPE) from the various combinations of PBL schemes and forcing datasets range between 28 % and 44 %. Whereas simulations with the MYJ scheme produce mixed (positive and negative) biases, all outputs for the MYNN3 scheme underestimate observations. Indeed, 143 Table 6.4 indicates that the MYJ scheme ameliorates overly dry conditions simulated by the MYNN3 scheme. The MYNN3 scheme biases decreases as one moves inland from the Kitimat station. Since the actual precipitation decrease away from the coast, the consistent drop in the MYNN3 biases indicates this scheme less represents the precipitation gradient from coast to inland, compared to the MYJ scheme (Fig. 6.3). Domain-wide precipitation outputs amongst the three datasets mostly differ by factors of 0.7–1.5 for simulations with theMYNN3 PBL scheme, with greater differences for those of the MYJ PBL scheme. Noteworthy is how the yield from the datasets are affected by the PBL schemes. The NAM_ANL and ERA5 outputs are quite comparable and greater(less) than the NARR output with the MYJ(MYNN3) PBL scheme. Remarkably, the greatest differences are over the southern and western portions of the domain where precipitation estimates differ by factors > 2. As these areas are coincident with the landfall of synoptic-scale storms from the Pacific Ocean, the spatial plots imply that the choice of PBL schemes is fundamental to frontal rain contributions of mid-latitude cyclones that are frequent in the region between October to March. With the MYNN3 scheme, diagnosis of boundary precipitation fields is perhaps suppressed, and the skill of individual datasets seems less consequential to model output with this choice of PBL scheme. Direct spatial comparison of precipitation from the two PBL schemes (Fig. 6.4) show much larger precipitation amounts for the MYJ scheme with NAM_ANL and ERA5. Differences in the capacity of either scheme to simulate precipitation from east-moving midlatitude cyclones are again alluded to in the NAM_ANL and ERA5 plots. In the case of NAM_ANL for instance, the MYJ scheme precipitation is about twice as much as 144 Figure 6.2 2017 monthly observed precipitation (grey bars), alongside percentage contribution of snow water equivalent. Table 6.4 PBIAS values for 2017 precipitation simulation at station locations. MPE is the mean of all PE per PBL scheme or dataset. Values in italics are the lowest biases of evaluations at each station. MYJ MYNN3 Kitimat Riverpark Lakelse Terrace Kitimat Riverpark Lakelse Terrace MPE NAM_ANL +21 +36 +38 +30 -58 -46 -40 -40 37 ERA5 -21 -12 -4 -7 -57 -46 -43 -33 28 NARR -55 -42 -36 -31 -53 -40 -36 -31 41 MPE 28 44 that from the MYNN3 scheme. For model forcing with NARR, the difference between PBL schemes is marginal (∼ 0.9–1.1 multiplicative ratio). Recall from Table 6.4 that the NAM_ANL and ERA5 outputs have more precipitation than NARR outputs. The resemblance of precipitation estimates from both PBL schemes for the NARR dataset, also evident in similarity of negative PBIAS values across stations, thus suggests persistent undercapture of the precipitation field over the valley with the NARR dataset. 145 146 Figure 6.3 Spatial plots of dataset-normalized annual precipitation estimates for simulations with MYJ PBL scheme (top) and MYNN3 PBL scheme (bottom). NAM_ANL/ERA5 means NAM_ANL output, normalized by ERA5 output, etc. Figure 6.4 Spatial plots of PBL scheme-normalized (MYJ/MYNN3) 2017 precipitation estimates for each dataset. For NARR it is MYNN3/MYJ. 6.3.2 Spatio-temporal verifications The spatial and temporal correspondence of precipitation estimates with observations are further assessed for outputs from the MYJ PBL scheme whose MPE is smaller than that 147 of the MYNN3 scheme. Additional information is gained from examining monthly totals and Fig. 6.5 shows that all three datasets represent the October peak precipitation to varying degrees; however, NARR does not depict the March and November accumulations that are also the months of peak contributions from snow (refer to Fig. 6.2). In essence, a main issue with the NARR output is poor attribution of snow amounts in the southern part of the valley — virtually all the snowfall at the Kitimat station is unaccounted in the NARR output. At the Kitimat station, less than 2 % of total precipitation for the NARR simulation is attributed to snowfall (Table 6.5), in contrast to 12 %, 8.5 % and 12.5 % for the station record, NAM_ANL and ERA5, respectively. The monthly NARR outputs however, are more positively correlated with station data than those of ERA5, perhaps due to it producing less precipitation. Overall the best correlations are for NAM_ANL, exceeding 0.70 at all four stations. Table 6.5 Contributions of snow water equivalent to total precipitation for observations (OBS) and those of simulations with various datasets and MYJ PBL scheme. Kitimat Terrace OBS NAM_ANL ERA5 NARR OBS NAM_ANL ERA5 NARR Snow water equivalent (mm) 284.4 246.4 240.4 123.0 142.6 127.4 1875.8 1103.9 1168.0 1543.8 1089.2 803.9 Total precipitation (mm) 2380.7 2890.7 % snow contribution 12.0 8.5 12.8 42.7 1.8 3.7 9.2 11.7 Whereas month-by-month correlations can evaluate temporal fitness, pairwise comparisons of seasonal contributions to annual precipitation can be more insightful of precip148 85.8 10.7 Figure 6.5 The 2017 monthly series of observations (OBS) at gauge stations, and simulations with the MYJ PBL scheme for different datasets. itation dynamics. This is because precipitation events are more frequent and intense at certain periods of the year, than others. For the purpose of evaluating model outputs, the year-long simulation is partitioned into four seasons namely spring (March–May), summer (June–August), autumn (September–November) and winter (December–February). Season-normalized proportions for all four seasons, at each station is presented in Fig. 6.6. Changes in accumulated precipitation across seasons are best emulated by NAM_ANL as it is often the closest to the measurement ratios. This correspondence is more for Kitimat and Riverpark, than for the other two stations that are further inland, thus demonstrating 149 the best representation of the seasonality of precipitation in the wetter, southern portion of the valley. Figure 6.6 Pairwise seasons-normalized 2017 precipitation for observations (OBS) at gauge locations, and simulations with MYJ PBL scheme and the various meteorological datasets. Analysis is extended to assessing the representativeness of inherent spatial variability of precipitation by tracking the proportionality of changes in model outputs. The stationnormalized plot of annual precipitation amounts for gauge measurements and model 150 outputs (Fig. 6.7) shows that spatial variability in the simulations is comparable to those of measurements for the most part. Over longer separation distances between any two gauges, modeled ratios tend towards being smaller than those of observations. The greater disparities for Kitimat/Lakelse and Kitimat/Terrace in particular, may imply that the rain shadow effect of coastal mountains that result in drier conditions in the northern part of the valley is less in the simulations than in reality. However, it should be noted that measurements are also uncertain. For instance, although 2017 precipitation records at the Lakelse station were complete, this station measures weekly accumulations, and measurement errors are difficult to identify at longer reporting frequencies. Further, all the stations are within the valley channel and precipitation undercatch, which can be substantial, particularly, for snowfall and in windy conditions, may have resulted in higher ratios than is correct. Nonetheless, the NARR ratios are consistently the most deviant from measurement data, suggesting that NARR precipitation magnitudes are less suited to downsaling in this region. 6.3.3 Predictive evaluations versus summary distributions Evaluations of the uncertainity in precipitation simulations are often geared towards improving the accuracy of short range forecasts, particularly for daily periods. Consequently, the use of conventional forecast indices has become widespread but such practice may not be revealing of the credibility of model outputs for quantitative applications. For the majority of precipitation events (days with accumulation ≥ 0.2 mm), forecast scores among the datasets are comparable (Table 6.6), de-emphasizing the amount deficit of the 151 Figure 6.7 Pairwise stations-normalized 2017 precipitation for observations (OBS) at gauge locations, and simulations with MYJ PBL scheme and the various meteorological datasets. NARR outputs in comparison to those of NAM_ANL and ERA5. Predictive outcomes for a smaller subset of precipitation events (daily amounts ≥ the mean for all days with precipitation) are more differentiated and show the NARR forecast performs worse than those of the other two datasets. However, performance indicators are not always consistent. As an example, whereas POD, ETS and FBI at the Kitimat station are more favorable for ERA5 than NARR, the latter has a lower false alarm ratio. Poorer forecast quality with smaller data size misrepresents the adequacy of all three datasets. Indeed, timing accuracies, exemplified by numerical forecast scores can be indecisive for choice of precipitation product where environmental consequence of absolute amounts is the end-purpose. For instance, the frequency of precipitation quantities ≥ the mean of daily events may be a small subset of all events, yet could be much impactful for the removal of pollutants from the atmosphere 152 to the surface. For this reason, quantitative evaluation metrics that are relevant to atmospheric deposition are more ideal. , ERA5 NARR NAM_ANL NAM_ANL 0.79 0.11 0.18 0.14 0.54 0.40 0.43 0.95 0.99 0.92 Riverpark 0.85 0.82 0.78 0.16 0.22 0.21 0.49 0.38 0.37 1.01 1.05 0.99 Terrace 0.84 0.75 0.78 0.22 0.26 0.25 0.43 0.31 0.35 1.07 1.01 1.03 Kitimat 0.77 0.48 0.20 0.32 0.41 0.33 0.48 0.27 0.14 1.13 0.81 0.30 Riverpark 0.81 0.54 0.32 0.43 0.42 0.43 0.42 0.30 0.19 1.42 0.97 0.56 Terrace 0.28 0.49 0.55 0.64 0.37 0.27 0.13 1.54 1.30 0.78 NARR NAM_ANL 0.85 0.82 ERA5 NARR Kitimat NARR Station ERA5 FBI (perfect =1) ERA5 Target POD (perfect =1) FAR (perfect =0) ETS (perfect =1) NAM_ANL Table 6.6 Predictive scores of daily precipitation for all events (≥ 0.2 mm) and events ≥ mean daily amounts. Mean daily amounts are 11.7, 9.6 and 6.9 mm at the Kitimat, Riverpark and Terrace stations, respectively. All daily events (amount ≥ 0.2mm) ≥ Mean daily quantity 0.78 0.59 Figures 6.8 and 6.9 depict the fitness of datasets based on event frequencies. At monitoring stations, daily events ≥ the mean daily precipitation account for no less than 25 % and 70 % frequency and amount, respectively, of all events (Fig. 6.8), with the NAM_ANL and ERA5 outputs reasonably matching these contributions. Fig. 6.9 also highlights the uncertainity of simulated precipitation. Apart from the 10th decile, the mean of each decile from NARR does not exceed the annual mean of all events days in Kitimat. This is unlike the other datasets where the means of the top 3-4 deciles are greater than the average of station data. The other stations equally depict poor performance for NARR. Considering 153 that the observation quantiles are bounded by ERA5 and NAM_ANL, it is concluded that both datasets provide more credible estimates of precipitation than NARR. Figure 6.8 Categorical evaluation of simulated precipitation amounts by various datasets with the MYJ PBL schemes with respect to station data (OBS) for all daily events (days with precipitation amounts ≥ 0.2 mm), and daily events ≥ the mean daily precipitation amount. The latter’s contributions in terms of frequency and amount are also plotted. 6.4 Discussion Across stations, the lowest biases in simulated annual precipitation ranged between 21 % underestimation to 3 % overestimation. This range is smaller than the -15 % to +25 % uncertainty band across seasons reported by Wong et al. (2017) for the Pacific Maritime 154 Figure 6.9 Cumulative mean distributions of simulated and observed daily precipitation at validation stations. The amount at each decile is the mean of all values within that decile. Dashed vertical line is mean precipitation of daily events. terrestrial ecozone of Canada. That study evaluated five gridded precipitation products against precipitation-gauge station data but, was based on a regridding to a coarser resolution (0.5 ◦ ) for all compared datasets. No intermediary regional climate model was used. Lilhare et al. (2019) reported annual precipitation biases ranging -8.2% to 13.5% for five hydroclimatic datasets over the Lower Nelson River Basin in the Canadian Shield. Their study was as well based on regridding to 10 km (∼ 0.1◦ ) spatial resolution with 155 bilinear interpolation (no dynamical downscaling), hence precluded the role of model physics. Datasets in that study mostly exhibited positive biases, unlike results in the present study. Aside from being over relatively flat terrain, Lilhare et al. (2019)’s study was for a sub-arctic continental climate with annual observed precipitation of ∼ 500 mm and peak precipitation during summer. Dissimilar geographies and methodologies could therefore be the principal reasons for differences in uncertainty ranges between this study and other Canadian investigation. Because model outputs over the TKV were from simulations at high horizontal resolution (∼ 0.01◦ ) with various combinations of meteorological forcing data and PBL schemes, the bias range in this study provides information about WRF competency to reproduce point observations of precipitation in a coastal valley region influenced by mid-latitude cyclones originating over the Pacific Ocean. With a sparse observation network and more dependency on parent model forecasts, uncertainties in physical parameterizations of atmospheric processes become prominent. Overall and irrespective of dataset, simulations using the MYNN3 PBL scheme exhibited larger and more negative biases than simulations using the MYJ scheme. WRF physics parameterizations are highly integrated, with PBL schemes providing the interface between cumulus, microphysics and radiation schemes on one hand, and surface-layer and land surface schemes on the other. While both tested PBL schemes used similar physics settings, significant portions of parent and intermediate grids, at which spatial scales convective eddies are parameterized, consisted of water (the Pacific Ocean). The MYNN3 scheme was designed to improve several aspects of earlier Mellor-Yamada schemes such as an insufficient growth of the convective boundary layer and underestimation of the 156 turbulent kinetic energy and the turbulent length scale (Nakanishi and Niino 2009), but such concerns are perhaps more applicable to warm, dry climates. Moreover, much of the precipitation in the TKV occurs during the months of October-March when east-going, low pressure systems and frontal passages dominate weather. Convective periods are infrequent and precipitation that may be produced by this process, such as during summer contributes little to total annual amount. Better model performance with the MYJ scheme is similar to findings by Evans et al. (2011) for the east coast of Australia that experiences heavy rain and strong onshore winds as occurs over the northern BC coast. In geographical areas with convective precipitation (e.g. Srinivas et al. 2018), MYNN-type schemes have been reported to outperform the MYJ scheme for intense events. In the present study, it is suspected that the lack of convection in the cold wet season suppresses precipitation simulation in the MYNN3 scheme much more than with the MYJ PBL scheme. Implicit therefore, is that MYJ PBL scheme is a better fit with the new Thompson microphysics scheme that was deployed for all model runs (Thompson et al. 2008), since non-convective precipitation in WRF is controlled by the microphysics scheme. At all sites, the driving dataset that gave results closest to observed annual total precipitation was ERA5, followed by NAM_ANL. With model runs using the MYJ PBL scheme, both datasets demonstrated a reasonable fit with precipitation distributions. Whereas ERA5 produced moderately low biased outputs for the most part, NAM_ANL outputs were biased high. Within the context of environmental change applications, such as modeling atmospheric deposition of air pollutants, these findings have important ramifications since precipitation fields are needed to drive relevant models. ERA5 has the poten- 157 tial of causing slightly less wet deposition than would occur since precipitation is biased low. In contrast, NAM_ANL would project much higher rates of wet deposition; possibly with greater spatial and temporal fitness than ERA5. Perhaps using both datasets for environmental modeling in the region would ensure that neither the quantitative accuracy of ERA5 nor the spatio-temporal performance of NAM_ANL would be lost. Such an approach will allow for projections that are realistic. The NARR dataset least quantified precipitation. The assessment of NARR as significantly dry-biased agrees with Wong et al. (2017) for their study over the whole of Canada. Nonetheless, in regions with much lower precipitation than northwest BC (e.g. Lilhare et al. 2019), NARR has been reported as slightly overestimating observations. Because of differences in nesting required to downscale each dataset within WRF, also considering the hydroclimatology of the TKV area, topographic resolution of the coarsest grid might also have influenced the performance of the meteorological datasets. The raw resolution of NAM_ANL is 12 km, and was more amenable to a downscaling ratio of 1:3 than the other datasets. This ratio, if used for ERA5 and NARR would have required more nested grids, with additional computational cost and perhaps more errors accumulating in the finest domain. While the ratio of 1:5 as used for the reanalyses datasets is permissible for precipitation downscaling (Liu et al. 2012), a ratio of 1:3 allowed for the use of higher-resolved terrain data for NAM_ANL. With a 9-km grid for the largest domain in the NAM_ANL simulations, topographic data resolution was 5-arc-minute and not 20-arc-minute that was used for the coarsest grids in ERA5/NARR simulations. For the intermediate and TKV domains, topographic resolution was the same for all three 158 datasets. Since the north BC coast has steep elevation gradients and complex topography, finer representation of surface terrain for the 9-km grid in NAM_ANL simulations may have contributed to it having less bias from observational data than NARR. However, as this study was not for an entirely continental setting, the benefit of finer-resolved topography could be small since a large portion of the coarsest grids and intermediate grids for all three datasets is water (the Pacific Ocean) and not land. Another plausible reason for differences in performance would be inherent quality of datasets. Better spatio-temporal correspondence of NAM_ANL with observations than for ERA5 and NARR was found despite that NAM_ANL is an analyses product, available at 6-hourly interval. Although improvements are being made, such as implementing a four-dimensional data assimilation system, and increasing grid and temporal resolutions (e.g. ERA5), reanalysis products are still much reliant on the number of observations that are ingested. Observation data sparsity particularly in remote regions translates to low quality of atmospheric products for these areas. Over northwestern Canada, there are not many weather stations—a status that impacts the quality of atmospheric data products for the region. All datasets would benefit from more quality control of assimilated data, including the treatment of missing observations. Dataset availability at higher temporal frequencies would also be helpful as this most likely contributed to better performance of ERA5 than NARR. Indeed overall poor performance with NARR, in spite of its specificity for North America, suggests a need for further enhancement of its value to the Pacific northwest. In NARR development, no data are assimilated over the oceans north of 43.5 ◦ N, and there has been non-assimilation of gauge data over Canada after 2002 (Mesinger 159 et al. 2006), ostensibly causing the large discrepancies with gauge records. This study compared model outputs using the same scale (hourly, 1-km2 grid-cells). An aspect seldom contemplated in many model evaluation studies is whether the non-equivalence of spatio-temporal resolutions of atmospheric datasets is addressed when they are compared to one another. Analyses and reanalyses products, and even gridded interpolated records are often directy compared against one another, despite that they are mainly available at different space and/or temporal resolutions. The downscaling approach in this study not only created a common spatio-temporal reference, but also used a fine horizontal resolution. Gauge measurements were considered point values valid over a limited area. Hence, 1-km2 grid-cells were used to generate reasonably discrete outputs for the TKV area, although expending significant computing resources. This recognition is particularly relevant when computational tools are intended for assessment of local environmental concerns, rather than for projections across continental/global climates. 6.5 Concluding remarks The urgency of addressing the environmental impact (e.g. atmospheric deposition of air pollutants) of industrial projects that are planned in the TKV required evaluating various choices for initial and boundary meteorological forcing, and PBL schemes in WRF for precipitation simulation, where due to few surface weather observations, gridded outputs will be input to other models. For combinations of three atmospheric datasets and two PBL schemes, large (> 40 %) underestimation of precipitation was found for simulations 160 involving NARR as forcing data, or MYNN3 as a PBL scheme. Model configuration with either NAM or ERA5 as atmospheric forcing data and MYJ as a PBL scheme performed better. With this arrangement, the uncertainty in annual precipitation amount ranged between 40 % overestimation and 20 % underestimation. The differences between model outputs not only inform the relative suitability of various modeling configurations to provide realistic precipitation estimates, but also highlight the importance of continuous improvement of individual components for optimal performance over Pacific Canada. This could be through greater observational data assimilation into raw products, dataset production at higher spatial and temporal resolutions, and verifying PBL schemes over inhomogeneous terrain. 161 Chapter Seven Modeling terrestrial ecosystems exposure to incremental smelter sulfur dioxide emissions and deposition in a complex coastal valley airshed Abstract Predicting the impact of air pollution from industries is important for developing management strategies under changing emissions. This chapter assessed the effect of increased SO2 emissions from a large aluminum smelter in the Terrace-Kitimat valley, a predominantly pristine airshed in northwestern British Columbia, Canada, on the natural environment using an atmospheric-chemistry model. CMAQ simulations with 1 km2 grid cells indicated at least 50 % increase in ambient SO2 concentrations due to rise from previously 27 tonnes day−1 smelter emissions rate to 42 tonnes day −1 maximum permis- 162 sible discharge. The risk of harm from acidifying emissions on flora, and on soils from new levels of SO2 exposure and deposition were evaluated based on critical level and critical acidity load values, respectively. Comparisons to baseline impacts estimated 50–88 % increase in aerial exceedance of limit for protection of lichen, and 37–67 % increase in spatial exceedance of threshold for protection of soils. Cumulatively, 16–18 km2 of plant habitat and 10–11 km2 of soil in an area contiguous with the smelter site will likely be damaged by its SO2 emission under the latest regulation. Whereas these projections are consistent with expectations of marginal spatial impact from tall stack emissions, it is imperative that soils are routinely sampled for phytotoxicity beyond the range outlined in this study. 7.1 Introduction Industrial air waste permitting by regulatory agencies often prioritize environment protection goals ahead of profit-making by businesses (Dietz 2003, Longley 2019). This is because many production processes are emission-intensive and their environmental impact may linger, long after operations have ceased. For example, it is common knowledge that some feedstock in metals manufacturing produce environmentally damaging emissions (Andrews and Lattanzio 2013, Cirtina et al. 2016, Habashi 2011). A precautionary approach to airshed management thus requires regulating emissions from major sources so that neither the resultant pollutant concentration in the atmosphere, nor deposition to ground receptors are occurring at levels that cause significant harm to specified sensitive components of an ecosystem. Critical levels which are exposure limits above which 163 adverse effects of ambient concentrations on key biota are observed, and critical loads which are deposition thresholds above which acidification and eutrophication will occur in the long term (1-100 years), have been proposed for a range of environments (Coordination Centre for Effects 2017, Josipovic et al. 2011, WHO 2000). Originally addressing issues pertaining to transboundary acidifying emissions impact on ecosystems, critical levels and critical loads have become pivotal to managing local air emissions including in Canada where they are being used to guide decisions concerning industrial development in wilderness areas. Chemical transport models are useful to determining compliance to critical levels and loads over geographical areas since they are able to simulate ambient concentrations and deposition of air pollutants at discrete locations far away from monitoring stations. Because they ingest meteorological and terrain data, in addition to implementing robust physical and chemical interactions between multiple pollutants, these tools provide a means to link spatial concentrations/depositions of pollutants to emissions (Brook et al. 2019, Kelly et al. 2018, Vivanco et al. 2018). This advantage is routinely exploited for indicating and projecting air pollution and acid deposition effects, including over complex environments, such as deep valleys (e.g. Kelly et al. 2018). In the past decade, air dispersion and deposition modeling in topographically-confined areas of British Columbia began placing more emphasis on effects of aggregate emissions from multiple sources (ESSA Technologies et al. 2014, Krzyzanowski 2010). Overly aggregated analyses however, may obscure contributions of large sources to potential concerns. Environmental impact attribution to industrial emissions is an integral part of social ac164 countability particularly when they are present near or within natural resource-based communities. This requirement becomes more compelling as pollutant emission levels change. Identifying exposure and deposition fields of terrestrial ecosystems in response to changes in emissions helps track the course of effects. It also facilitates designating environmental responsibility to polluters. Further, it can provide valuable information on solutions, for instance, whether mitigations should be more at the level of sources or at receptors. Therefore, while cumulative impacts assessment (Krzyzanowski 2010, Pickard et al. 2019) remains important, specifying the impact of emissions change of single, large sources is also pivotal in areas that host such sources. This chapter assesses the magnitude and extent of the effect of recent aluminum smelter emissions change (refer to Fig. 5.1 for its position), on the land ecosystem in the TKV. The CMAQ model is used to simulate air concentrations of SO2 , and deposition of sulfur (S). Attention is primarily on comparisons of model outputs to critical levels of exposure of sensitive organisms to SO2 , and to critical loads of S, as an indicator of potential soil acidification. In the next section, existing and incremental industrial emissions, as well as the modeling approach are described. In Section 8.3, modeling results, including changes in pollutant concentrations from baselines, and exceedance of natural resource protection limits are presented. Section 8.4 provides a discussion of the results while section 8.5 concludes the chapter. 165 7.2 Modeling framework and procedure The WRF-SMOKE-CMAQ modeling system consists of three main steps namely meteorological simulation, emissions modeling/ processing and air chemistry/deposition simulation. Year 2017 meteorological forcing fields at 1 km horizontal resolution over the TKV and surrounding areas were generated (see Chapter 6) from two datasets: the North American Mesoscale Analyses (NAM_ANL) and ERA5. These were used to provide reasonable precipitation estimates. A similar approach to emissions modeling with the SMOKE tool described in Chapter 3 was utilized for emissions processing, which in this study, was to generate two sets of gridded emissions for each meteorological input file. One relates to pre-existing emissions (that of ERA5 for emissions prior to aluminum smelter upgrade, or the pre-modernization discharge limit was already derived in Chapter 3), while the other was for smelter emissions up to maximum, 42 tonnes day−1 , SO2 discharge limit (post-modernization period). Emissions from all existing sources in 2017 in the airshed, including for the boundary file at 3 km grid spacing (the grid size of the intermediate domain of NAM_ANL input meteorology (Table 7.1) were from the BC portion of Canada-wide, SMOKE-ready inventory. To account for incremental ship-related emissions, the increase in quantity was based on scaling from total estimates (ESSA Technologies et al. 2014) in proportion to the projected movement of an additional 23 cargo vessels for increased export of finished products and import of raw materials. Actual allocation in space was according to mode of vessel and tugboat activity, in addition to consideration of regulations on ship fuels in North 166 American waters, and pollution emission factors of ocean-going vessels (USEPA 2009). For all pollutants including SO 2 , emissions from marine vessels traveling in the Douglas Channel were modeled as 40 waterway positions from the marine terminal next to the smelter down to the edge of the domain. Emissions at berth was fixed at half the total of underway releases (that is, 1/3 of all maritime transportation related-emissions). The spacing of positions was between 1–1.4 km with few clustered nearer the marine terminal to imitate higher traffic. All emissions from the smelter site, at the marine terminal, and for travel of vessels in Channel waters, were modeled as elevated point sources using discharge parameters indicated in regulatory/government publications (Environmental Appeal Board 2015, ESSA Technologies et al. 2014). The final step was running CMAQ in a two-sequence nesting mode: the first to derive boundary concentrations and the second, to generate air concentrations and deposition fluxes over the TKV area. For each meteorology forcing, runs with, and without emissions changes were performed for the period between 20th December 2016 to 2nd January 2018, retaining only outputs for 2017 for the 1-km grids. Hourly concentrations of SO2 were retrieved. Total depositions (wet plus dry deposition) of S were calculated using in-built formulas that add-up relevant species. Chapters 3 and 5 that reported satisfactory modeling of key acidifying air pollutants, and acidic wet depositions in the valley, respectively, can be consulted for baseline concentrations and deposition fluxes with ERA5. 167 Table 7.1 Meteorological and emissions domain sizes and attributes for SMOKECMAQ modeling with ERA5 and NAM_ANL data. Meteorological simulations were performed using the MYJ PBL scheme. Type Domains (dimensions) ERA5 NAM_ANL Parent grid size (km) 25 9 Grid (nesting) ratio 1:5 1:3 Parent (W-E × N-S) 100 x 100 100 x 100 Parent grid center 54.200 °N, 128.600 °W 54.200 °N, 128.600 °W Meteorology 1st nested (W-E × N-S) 121 x 121 121 x 121 1st nested grid center 53.850 °N, 128.795°W 54.158 °N, 128.671 °W 2nd nested (W-E × N-S) 101 x 121 100 x 121 2nd nested grid center 54.223°N, 128.640°W 54.213 °N, 128.695 °W Grid centers Emissions 7.3 54.200 °N, 128.600°W Boundary grid (W-E × N-S) 40 x 60 60 x 60 TKV grid (W-E × N-S) 36 x 106 36 x 106 Results and analyses 7.3.1 Baseline SO2 levels and changes For assessing changes in ecosystem exposure to SO2 , baseline pollutant levels are required and Fig. 7.1a shows modeled ambient concentrations for smelter SO2 emissions at the previous limit of 27 tonnes day−1 . Spatial plots are for simulations using NAM_ANL and ERA5 as sources of meteorological data, respectively. Except around the smelter site, 168 ambient SO2 levels are low. Highest annual average concentrations are within a plume directly north of the smelter site, but beyond 15 km northward, concentrations fall to less than 1 ppb, and less than 0.5 ppb by the halfway distance between Kitimat and Terrace. A southward plume barely exists, which could be expected since SO2 would readily be diluted by more frequent and stronger winds around the shoreline than encountered within the valley. The ERA5 simulation generates slightly higher concentrations than that of NAM_ANL at longer distance away from the smelter. NAM_ANL simulates more precipitation that may wash-out more pollutants, hence lesser amounts that would linger in the atmosphere. Nonetheless, both simulations are quite consistent in their indication of pollutant patterns, namely, a small area of intense concentration (> 10 ppb), and onshore emissions advection and confinement by valley’s west side-walls. Overall, modeled annual SO2 concentration ranged 0–23 ppb. Emissions alter pollutant baseline and Fig.7.1b shows the percentage increase in ambient SO2 due to smelter SO2 discharge permit amendment from 27 to 42 tonnes day−1 . It is evident that there is substantial rise in relative terms in view of at least 50 % increase over much of the valley area, although this in part would be due to pristine settings and preexisting low levels. Consequently, ambient concentration sensitivity to emissions change is high at the present stage of industrial development in the valley. The greatest concentrations changes are near the smelter source, which rise by as much as 100 % or more. Apart from slight differences in percentage rise over the middle and northern sections of the valley, spatial SO2 concentration gradients for ERA5 and NAM_ANL simulations are quite similar. For both, the relative rise in ambient SO2 extend throughout the valley area, 169 even though other anthropogenic emissions sources (e.g area sources) were unchanged. This demonstrates how important SO2 emissions from a single facility is at present in the airshed. Figure 7.1 Average annual SO2 levels for (a) ambient concentrations at smelter’s pre-modernization limit (b) relative increase in ambient concentration for discharge at present maximum threshold. 7.3.2 Critical level exceedances and mapping While relative changes in pollutant levels are useful, absolute concentrations are important for quantifying the magnitude and intensity of exposure of environmental compo170 nents to ambient pollution. Persistent SO2 exposure of agricultural crops, and vegetation is widely associated with decline in farm yield and forest health, respectively. For this reason, modeled SO2 concentrations for incremental emissions from the modernized smelter and allied activity are referenced to critical levels (CLv) outlined by the World Health Organization (WHO 2000) for floral communities. This allows identifying and mapping the spatial extent of excessive SO2 pollution that can be detrimental to vegetation. Because accumulated temperature sum above +5 ◦ C is > 1000 ◦ C·days per year in the TKV, the SO2 concentrations thresholds of 7.2 ppb for protecting forests and natural vegetation, and 3.6 ppb for sensitive lichen were deemed applicable CLv. Exceedance is the concentration of modeled SO2 above the CLv and spatial plots (for lichen alone) are displayed in Fig. 7.2. As expected, SO2 CLv exceedances occur around the smelter location. For vegetation and forest (Table 7.2), aerial SO2 exceedance is estimated to enclose an additional 2 km2 from the permitted maximum emissions during the pre-modernization period, alongside 1.4–1.6 ppb more SO2 exposure, or roughly a 20 % increase in intensity over the original affected areas. For lichen, an additional 6–7 km2 spatial exceedance area is modeled, meaning a 50–88 % increase in affected area from the baseline. Exceedances in the original areas are also estimated to increase by 1.3–1.4 ppb or about 30% on average. Westward, exceedance spread appears curtailed by bounding valley walls which is quite close to the smelter site. Instead, new exceedance areas are an elliptical area of elevated SO2 plume, spanning 3 km south (along the shoreline) and 5 km north of the aluminum smelter. There is some expansion eastward into lands across Kitimat River, although the general area on the east side of the Douglas Channel is unaffected. In fact, even for pre-modernization 171 Figure 7.2 SO2 spatial exceedance for lichen in the Kitimat area at smelter’s premodernization emissions limit (27 tonnes day−1 ) and maximum threshold (42 tonnes day−1 ). Exceedance is modeled SO2 minus CLv. SO2 emissions limit, there are no exceedances elsewhere in the domain, alluding to little impact of other pollutant sources (e.g residential wood burning, transportation) to ambient SO2 levels. Thus SO2 exceedance in the TKV at present primarily depends on the quantity of smelter SO2 emissions. 7.3.3 Critical load of acidity exceedances and mapping Annual wet, and dry S deposition output from CMAQ are also totaled for the two periods of smelter emissions, for calculating critical load of acidity (CLA) exceedances. Because CLA relate more to effects of S deposition on ecosystem structure and functioning over 172 Table 7.2 Estimated aerial SO2 exceedance of critical levels of vegetation and lichen exposures due to smelter emissions changes using ERA5 and NAM_ANL datasets. Exceedance is modeled SO2 minus CLv. Lichen and Forest Vegetation NAM_ANL 3 7.4 2 0.6 8.8 5 5.5 ERA5 3 8.9 2 0.3 10.5 5 6.4 NAM_ANL 9 4.4 7 0.8 5.8 16 3.6 ERA5 12 3.9 6 0.7 5.2 18 3.7 periods of decades, they indicate the state of existing or required environmental protection in the long-term. In this regard, the critical load that is reported for the area is used. The average CLA for terrestrial ecosystems in the TKV estimated by Williston et al. (2016) from steady-state mass balance is equivalent to 29 kg S ha−1 yr−1 . The mass balance models used as critical limits, base cation to aluminum ratios (Bc:Al) of 1, and 6 as critical limits for the predominant coniferous forests growing on mineral soils (65 % of the area), and deciduous forest tree species, respectively (ESSA Technologies et al. 2014, Williston et al. 2016). Their calculation for the TKV took consideration of mixed vegetation at 1 km2 resolution further facilitating usage in the present study. This CLA value is subtracted from modeled S deposition fields, thereby providing spatial extent of exceedances (Fig. 173 for cumulative area (ppb) Average exceedance area (km2 ) Cumulative exceedance in previous area (ppb) Average exceedance in additional area (ppb) Average exceedance area (km2 ) exceedance (ppb) Average Total area (km2 ) forcing Post modernization Additional exceedance Atmospheric Pre-modernization 7.3). Area-weighted exceedance magnitudes are also calculated (Table 7.3). Figure 7.3 Modeled exceedances of CLA from SO2 emissions of aluminum smelter in Kitimat at pre- and post-modernization rates with ERA5 and NAM_ANL datasets. Exceedance is modeled sulfur deposition minus CLA Patterns of CLA exceedance for the two emission periods are broadly similar to ambient SO2 exceedance of critical limits in Fig. 8.3. However, modeled CLA exceedance areas are smaller. In particular, the exceedance area does not extend as far south as seen for the SO2 exceedance. It is unclear why this is so, since frequent and large precipitation amounts in the coastal part of the valley should cause more gaseous SO2 to be removed from the atmosphere. This probably is an artefact of contouring different quantities (S versus SO2 ). Nonetheless, CLA exceedance is projected to broaden by 3–4 km2 if the smelter emits SO2 174 Table 7.3 Estimated exceedances of CLA from SO2 emissions of aluminum smelter in Kitimat at pre- and post-modernization rates with ERA5 and NAM_ANL datasets. for cumulative area (ha−1 yr−1 ) Average exceedance area (km2 ) Cumulative exceedance in previous area (ha−1 yr−1 ) Average exceedance in additional area (ha −1 yr−1 ) area (km2 ) Average exceedance Post modernization Additional exceedance exceedance (ha−1 yr−1 ) Average Total area (km2 ) Pre-modernization NAM_ANL 6 41.1 4 8.7 49.8 10 33.5 ERA5 8 37.9 3 10.6 48.0 11 37.8 at its increased level. The average exceedances of S deposition for the cumulative and the pre-modernization areas are quite comparable. But as Table 7.3 also indicates, and analogous to SO2 critical limits, exceedance in the original areas will amplify. From model outputs, exceedent deposition will increase by 8.7–10.1 kg S ha−1 yr−1 (∼ 21–27 % rise) in the previous impact area. These increments between emission amounts point to greater aerial impact of cumulative sulfur deposition nearer the smelter source. 7.4 Discussion CMAQ model outputs found that a 56 % smelter source SO2 increase translated to ≥ 50% increase in ambient levels over the valley. This similarity is expected, since ignor175 ing transformation and removal mechanisms, ambient concentrations are linearly related to emission amount, and the smelter source is the dominant source in the airshed. The highest modeled annual concentrations along a 70-km north-south valley transect following emissions increase (Fig. 7.4) is about 1 ppb but it is below this value for nearly the whole valley length, comparable to levels at remote locations (> 100 km from major anthropogenic sources) in Western Canada (e.g. Hsu 2013). A greater increase in ambient levels over the valley than over surrounding ridges attested to the role of steep valley topography in limiting the smelter plume dispersal; however, concentration changes also depend on the method of discharge. At the pre-modernization SO2 rate of 27 tonnes day−1 in 2017, emissions at the smelter were collected and vented as pre-heated effluent from tall stacks. Consequently, SO2 is dispersed over a wide area which perhaps would not have been the case if source was at ground level. The attenuation of SO2 exposure risk near the smelter is thus enhanced by long-range transport and deposition. Nonetheless, the relative rise in ambient SO2 at as far as Terrace not only emphasize the role of valley atmosphere in redistributing industrial emissions but also the potential of long-range effects from such sources. Maximum permissible smelter emission was projected to yield SO2 concentrations 3.6–3.7 ppb above thresholds that are dangerous to sensitive lichens, in an area 16–18 km2 . At the same time, the cumulative area for vegetation that is exposed to SO2 was 5 km2 , with an average of 5.5–6.4 ppb above the critical level. Because the critical limit for lichen was lower than for forest vegetation, non-exceedance for lichens assures the spatial range of short-term protection of plants from ambient SO2 pollution. The expanded effects 176 Figure 7.4 Annual mean concentrations from smelter SO2 source at maximum emissions rate of 42 tonnes day−1 along a north-south transect through the valley (right). area relative to previous exceedance however, suggests increased threat to their abundance. Regarded endemic to the Kitimat foreshore are rare lichens such as Cryptic paw (Nephroma occultum), Norman (Tholurna dissimilis), and Old-growth specklebelly (Pseudocyphellaria rainierensis) (Besse 2010, Committee on the Status of Endangered Wildlife in Canada 2006, United States Department of Agriculture 2013), and it is likely that they will be extirpated in nascent exceedance areas. Such prospect is worrisome considering the limited natural resilience for some species. Nephroma occultum for instance has poor dispersal efficiency and is easily displaced by competitors (Committee on the Status of Endangered Wildlife in Canada 2006). Hence, excessive SO2 concentration could pose an extra burden for their survival in those areas of exceedance. Still, on evidence of modeled exceedance, it appeared that pollution risk on lichen growth for areas east of Minette Bay is low and the advantage of little human disturbance for the most part might sustain local presence. At this time however, it is not known how lichen distribution has changed 177 with industrial SO2 emission rates in the valley because no relevant field survey has been conducted since more than 30 years. In contrast, vegetation has been inspected for foliar injury on a biennial basis since 1970. Inspections are usually along established transects, although conditions at several remote sites have also been captured. The commonest observations in the vicinity of the smelter have been leaf chlorosis, leaf notching and tissue necrosis but symptoms to date have mingled with pests and pathogens attacks, as well as exposure to historically significant hydrogen-fluoride emissions (Stantec 2015). Tissue concentration of sulfur in westernhemlock needles have also been analyzed at the same time interval as visual inspections. Foliar sulfur content has slightly risen following increase in SO2 smelter emissions above the previous 27 tonnes day−1 limit (Rio Tinto 2018). Since SO2 concentrations generally attenuate with distance from the smelter source, the nearest sampling sites would be sulfurous. Specifically, 5 out of 8 vegetation sample locations having the highest sulfur content in hemlock foliage in the Stantec report are within the gridded exceedance of SO2 exposure at pre-modernization emission limits. It thus can be expected that foliar S will rise in vegetation that falls under the new SO2 exceedance area. Acidification of terrestrial ecosystems, assuming the smelter emits SO2 at maximum permissible amount was estimated to comprise an area of 10–11 km2 centered around the facility, with mean exceedance of 33.5–37.8 kg S ha−1 yr−1 . Estimated spatial exceedance was dependent on a generalized critical load of acidity for the area which in turn relied on calculations of base cation weathering rates, base cation uptake (removal) in harvested biomass, critical acid neutralizing capacity, and several other predictor variables (ESSA 178 Technologies et al. 2014, Williston et al. 2016). Different methodological approaches in the derivation of critical loads can result in different conclusions about soil acidification status. For example, recent analyses across sites in the TKV (Levasseur et al. 2020) that combined mineralogy-based and particle-size properties adapted to local factors for calculating base cation weathering rates, produced high critical loads and no exceedance of S deposition. The critical load as developed by Williston et al. (2016) and used in the present study derives from a widely-applied texture-based function (Sverdrup and Warfvinge 1995) for estimating surface area and weathering rates (United Nations Economic Commission for Europe 2017, Whitfield et al. 2018). Therefore, even as acidification due to new smelter SO2 emission is projected to increase over a limited area, deposition estimates in this study represent a worst-case prediction. Although the approach to emissions release such as using tall stacks, high exit velocity, etc. alleviates SO2 vegetal exposure in the immediate surrounding, surveillance for indirect pollution is still advisable. Observations (Stantec 2015) have found poor growth of vegetation in proximity to the smelter site (within 1 to 2 km) suggesting plant toxicity effect of soil acidification. This is unsurprising considering that the smelter has been in continuous operation for decades; however, such outcomes can persist or become severe with future operations. Indeed, that modeled average exceedance of S deposition was more than 100 % the critical load, and extended as far north as 5 km from the facility, foretells further soil deterioration. Whether significant ecological values will be lost if projections are borne out, is unclear since the spatial footprint of adverse impact would mainly be on lands pre-zoned for industrial use in Kitimat (see Fig. 1.3). Nonetheless, on 179 account of the possibility of about 1225 tonnes excess S in 35 years of future aluminum production being received over an area that can contain 333,000–400,000 mature trees (see Appendix I), remedial measures may well be sought much in advance of the smelter’s decommissioning. 7.5 Concluding remarks The assessment of the effect of a 56 % increase in maximum SO2 emissions from a large aluminum smelter in Kitimat, with an atmospheric-chemistry model projected 50–88 % increase in aerial exceedance of limit for protection of lichen and 37–67 % increase in spatial exceedance of threshold for protection of soils. In total, 16–18 km2 of plant habitat and 10–11 km2 of soil in an area contiguous with the smelter site will likely be damaged by its SO2 emissions at the maximum permitted levels. Smelter emission of nitrogen oxides (NOx ), though included in the modeling are minor, and were not part of the analyses. However, several other industrial projects (e.g. liquefied natural gas facilities) are being developed in the valley, which when operational, will emit large amounts of SO2 and NO x . These would add to total pollutant emissions and the next chapter will examine their potential to further expand aerial exceedance of critical levels and critical loads. 180 Chapter Eight Quantifying incremental and cumulative terrestrial ecosystems impacts of NOx and SO2 emissions from LNG operations in the Terrace-Kitimat valley of northwestern British Columbia Abstract Natural resource development projects in British Columbia, Canada are subject to impact assessments to protect health and the environment. This study modeled incremental and cumulative terrestrial ecosystem effects of NOx and SO 2 emissions from two planned LNG facilities in the Terrace-Kitimat valley, a predominantly pristine coastal airshed in northwestern British Columbia. Emissions from these projects will cause at least 50 % 181 and 150 % rise in ambient SO2 and NO x , respectively; however only an additional 4 km2 will be exposed to SO2 concentrations that is directly harmful to vegetation. Cumulative NOx concentrations are expected to remain below harmful levels. For indirect vegetation impacts via soil pollution, limited areal exceedance (≤ 5 km2 ) of nitrogen deposition will barely be altered (0–1 km2 increase) by new emissions. Incremental SO2 discharge translated to 3 km2 extra exceedance of critical load of acidity and 13–14 km2 in total. Total area-weighted sulfur exceedance ranged 29.7–35.0 kg ha−1 yr−1 . Setting target loads, routine liming and ample distance offset of emission areas from sensitive habitats are measures that can mitigate soil acidification and irreversible changes in vegetation composition. 8.1 Introduction The quest by the Canadian province of British Columbia (BC) to be a major player in the global energy market by adding value to natural resources has meant increased commercialization of the downstream petroleum sector. One strategy is to site liquefied natural gas (LNG) processing and storage facilities at coastal locations for export overseas. LNG is natural gas that has been condensed into liquid at approximately -162 ◦ C, near atmospheric pressure, for easy and safe maritime transport (Center for Liquefied Natural Gas 2020). LNG has fewer emissions than other conventional fuels such as diesel and gasoline, also serving a viable alternate energy source for industrial, domestic and commercial uses (Center for Liquefied Natural Gas 2020, Smajla et al. 2019). Its production for export is hinged on abundant shale gas reserves in the northeastern part of BC (BC Oil and Gas 182 Commission 2020), concomitant with high demand in the Asia-Pacific region (Aguilera et al. 2014, BP 2020). Investment in the LNG business is expected to boost government revenue and provide more jobs. Potential monetary benefits notwithstanding, the location of LNG projects in pristine areas has caused concerns about adverse environmental effects. Other anxieties related to the LNG industry such as opposition from some First Nations communities to the construction of pipelines (The Interior News 2020) have also been reported. Construction of new LNG infrastructure in BC falls under large-scale projects that are subject to assessments and reviews by the government and the public through an environmental regulatory and permitting system. This process includes the evaluation of likely effects on key-value components such as the atmosphere. LNG projects when operational, emit sulfur dioxide (SO2 ), nitrogen oxide (NO), nitrogen dioxide (NO2 ), and other air pollutants. Exposure of plants to high levels of SO2 and NO x (NO2 + NO) can cause direct injury to leaves (Gheorge and Ion 2011), reduce crop yield through disruption of photosynthetic ability (Sun et al. 2016) and increase plant susceptibility to other environmental stresses, e.g. pests (Blande et al. 2014, Vacek and Matějka 2010). Further, SO2 and NO2 have acidifying properties which alongside other sulfur- and nitrogen-bearing compounds, can be deposited to soils. Excessive sulfur and nitrogen in base-poor soils can make them more acidic, causing leaching of nutrients and decline in forest productivity (Chumanová-Vávrová et al. 2015, Duan et al. 2016). Surplus soil nitrogen can cause certain plants to proliferate at the expense of others, leading to the disappearance of native flora (Bobbink et al. 2010). While injury to plants from acute and chronic exposure 183 to acidifying air pollutants may be detected through field inspections, direct observation of future scenarios is not possible. Addressing potential atmospheric pollution is complicated because the effects on ecological systems are lagged, making it difficult to ascertain the spatial progression of damage. In BC, regulatory atmospheric effects assessment of large development projects is conducted with the Lagrangian CALPUFF dispersion model (Sakiyama 2015). Although accepted as a screening modeling technique for long-range (≥ 50 km) transport, CALPUFF is typically not recommended for setting progress goals for multiple sources of very reactive pollutants (USEPA 2015, 2017). CALPUFF uses fixed, uniform concentrations of important oxidants such as ozone and neutralizing agents such as ammonia, with an overly simplified representation of secondary-formed chemical species (USEPA 2015). Methodological approaches so far are also less attentive to incremental impacts of industrial activities. Overly aggregated air dispersion modeling, whereby the contribution of an industrial sector is indistinct from pre-existing baselines (e.g ESSA Technologies et al. 2014, 2016), blur their indication of culpability for potential ecological harm. The foregoing not only suggest the need for procedures that are more consonant with environmental accounting but also, to using tools that are reflective of the state-of-the-science in chemical dispersion phenomena. The goal of this chapter is to predict atmospheric deposition of acidifying pollutants from LNG operations in Kitimat, British Columbia, Canada. It uses the CMAQ model (https://www.epa.gov/cmaq) to simulate how terrestrial ecosystems exposure to acidifying pollutants may evolve as precursor emissions occur. This study therefore addresses 184 the following questions: 1.) To what additional extent will vegetation be exposed to harmful NOx and SO2 concentrations from LNG projects? 2.) How much direct vegetation exposure to harmful NO x and SO2 concentrations will result from aggregate industrial emissions? 3.) By how much will soil nitrogen enrichment and acidification change as a result of the LNG industry? 4.) What aerial exceedances of critical loads of nitrogen and sulfur deposition will arise from aggregate industrial emissions? The remainder of this chapter is organized as follows. In section 8.2 , the numerical modeling approach, together with the processing and gridding of anticipated LNG pollutant emissions, are described. Analysis of model outputs, in the context of incremental and aggregate exceedance of wellness thresholds of key forest components are presented in section 8.3. A discussion of land ecosystem modeling and management implications appears in section 8.4. Section 8.5 concludes the chapter. 8.2 LNG emissions and numerical modeling set-up Deposition modeling with WRF-SMOKE-CMAQ, forced by separate meteorological datasets (NAM_ANL and ERA5) follows the same procedure described in Chapter 7. The difference is in the processing of emissions from LNG operations for simulations with, and without their inclusion. Two LNG projects whose realization are more proximate in time than all others were considered. These are LNG Canada (currently under construction) and Kitimat LNG (currently waiting a final investment decision) (Fig 8.1). Emission estimates of major air contaminants (NOx , SO 2 , VOC and PM10 ) for the two projects relied on information from proponents’ descriptions and scoping reports (ESSA Technologies et al. 185 2014, Impact Assessment Agency of Canada 2019). Emissions were modeled under three major categories: sources at production sites; hoteling carriers at terminals; and LNG carriers in transit (Fig 8.2). Combined pollutant emissions at main production sites of LNG Canada and Kitimat LNG was proportioned based on tentative export capacities. Emissions at terminals were estimated based on arrivals and departures of 350 and 250 LNG carriers (including assist tugs) annually for LNG Canada and Kitimat LNG, respectively. The same expected maritime traffic was used to project emissions for carriers in transit. For both projects, off-berth emissions were modeled as a series of point sources 1-1.4 km apart (the smallest mean distance between adjacent emission grid cells) along the Douglas Channel up to the boundary of the modeling domain. While the number of discrete points was 43 for LNG Canada, that of Kitimat LNG was 29; the difference being due to the location of their respective terminals. These points were slightly more clustered nearer the terminal locations to mimic higher maritime traffic emissions on getting closer to them. All LNG industry-related emissions were processed as elevated point sources. Stack parameters were obtained from scoping reports or analogous facilities. Following emissions processing, CMAQ version 5.2 was run to generate air concentrations of NOx , SO2 and deposition fluxes. Simulations for contributions of LNG operations followed a brute-force approach, that is, with, and without LNG emissions. Model runs were for the period between 20th December 2016 to 2nd January 2018, retaining only outputs for 2017 for the 1-km grids. Hourly concentrations of NOx , and SO2 were retrieved. Total depositions (wet plus dry deposition) of N and S were calculated using in-built formulas in CMAQ that sum appropriate species. 186 Figure 8.1 Locations of two LNG projects in the Kitimat area. Blue markers are export terminal sites, separate from LNG production sites (red markers). 187 Figure 8.2 Annual emission estimates of major air pollutants from proposed LNG projects in Kitimat, including on-way shipping along the Douglas Channel. 8.3 Results and analyses 8.3.1 Relative changes in ambient NO x and SO 2 and exceedance of critical levels Chemical dispersion and transport affect residual pollutant quantities and Fig. 8.3 shows modeled changes in ambient concentrations of acidifying gases due to LNG emissions. 188 For NOx , a worse-case projection (ERA5) is at least 150 % increase from the industrial baseline over much of the valley area although this in part would be due to largely to pristine settings and pre-existing low levels. For example, within the valley, precisely over Terrace that is more urbanized than Kitimat, increase is negligible (< 20%). Compared to NOx , the change in ambient SO 2 is lower (∼ 30–60 % increase) mainly because of the existing emissions from an aluminum smelter in Kitimat. The greatest SO2 concentrations changes are projected to be over small areas just outside the valley—around Kitamaat Village where levels would rise by as much as 100 %. These areas are contiguous with allied LNG activity such as shipping. Comparison of modeled annual SO2 concentrations to the critical level (CLv : 3.6 ppb annual average) of protection of lichen (WHO 2000) which is the most sensitive of floral communities (Fig. 8.3, Table 8.1), shows increase in aerial exceedance by 4 km2 (22–25 % increase). The exceedance increment appears along a north-south axis that marginally elongates the original elliptical area, which is to be expected since sideways spread of precursor plume by wind, particularly to the west is curtailed by steep topography. Although industrial source plumes are additive, excedent SO2 does not encompass the Kitimat LNG plant site, on account of smaller emitted quantity as well as dispersal by stronger winds in the maritime channel. However, the intensity of SO2 exposure is amplified in the near field of Canada LNG source. Whereas area-weighted exceedance with and without LNG emissions are comparable, there is 0.6–0.7 ppb more SO2 exposure over the original exceedance areas with the LNG emissions. Despite modeled increase, NOx ambient concentrations inclusive of LNG emissions are 189 Figure 8.3 NOx and SO 2 concentration changes relative to levels without contributions from LNG industry using the CMAQ modeling system driven by two different meteorological datasets. well below the critical level (15.6 ppb) for protection of vegetation (not shown). Highest modeled annual NO x concentrations in the vicinity of proposed LNG industries was roughly 5 ppb. Top concentrations were also limited to few grid cells, pointing to anthropogenic release locations as the commonest sources of non-negligible ambient NOx . Apart from the emitted quantity, shorter atmospheric lifetime of NOx compared to SO2 190 Figure 8.4 Exceedance (modeled concentration minus critical level) of lichen exposure to SO2 with, and without contributions from LNG industry emissions. (Seinfeld and Pandis 2016), whereby the former undergo several chemical reactions probably contributes to NO x non–exceedance. NOx concentration is mediated by natural ozone formation and destruction processes, including the formation of several atmospheric radicals. In view that the highest simulated concentrations were an order of magnitude less than CLv throughout the modeling domain, direct harm to vegetation arising from exposure to NOx concentration should be of low concern. 191 Table 8.1 Estimated exceedances (Modeled SO2 minus 3.6 ppb) of lichen exposure with and without LNG emissions using ERA5 and NAM_ANL datasets. The cumulative area is previous area plus new area NAM_ANL 16 3.6 4 0.7 4.2 20 3.5 ERA5 18 3.7 4 0.8 4.7 22 3.7 for cumulative area (ppb) Average exceedance area (km2 ) Cumulative exceedance in previous area (ppb) Average exceedance in additional area (ppb) area (km2 ) Average exceedance With LNG emissions Additional exceedance exceedance (ppb) Average Total area (km2 ) Without LNG 8.3.2 Relative changes in nitrogen and sulfur deposition and exceedance of critical loads A portion of emitted NO x and SO2 will be deposited while being transported in the valley, and Fig. 8.5 shows the relative changes in N and S deposition to the surface that is anticipated with LNG emissions. For both, the increase in deposition from pre-existing deposition amount is modest (20–50 % rise) within the valley but less over high elevations. This distinction in space, which is analogous to patterns for precursor concentrations in Fig. 8.3, demonstrates the constraining of pollutant transport by valley side-walls. Emitted SO2 will be deposited more in the valley channel than over the ridges. For S deposition, only in the vicinity Kitimat LNG facility would the relative increase be large (≥ 150 %), consistent with completely new industrial development in the area. Beyond this location 192 to the sout, deposition fluxes would barely be altered, suggesting that spatial deposition would also be controled by onshore winds that weaken as they approach the valley. Figure 8.5 Nitrogen (N) and sulfur (S) deposition changes relative to the loads without contributions from LNG industry emissions Evaluation of exceedance of critical load of nutrient nitrogen (CLNnut ) indicates that LNG NOx emissions would scarcely result in new exceedances, even around project sites. The CLNnut is an empirical deposition of 4 kg N ha−1 yr−1 below which semi-natural ter- 193 restrial habitats are protected. In the present, within ≤ 5 km2 combined area around Terrace is the CLNnut exceeded, with an additional 0-1 km2 projected for new emissions (Table 8.2). LNG operations probably would lead to a small increase in the magnitude of exceedance (∼ 0.2–0.3 kg ha−1 yr−1 more N overall). Qualitatively, excess N deposition from LNG emissions project low impact, not only on account of negligible affected area but also from an understanding that aerial exceedance, if any, would bear mostly on lands already under urban use. Nitrogen deposition in the valley at present, pertains to the Terrace area, and stems from very local sources. The major nitrogen-bearing emissions around Terrace are from road and rail transportation, rather than industrial sources. Table 8.2 Estimated exceedances (modeled nitrogen deposition minus 4 kg ha−1 yr−1 ) of critical load of soil nutrient nitrogen with, and without LNG emissions using ERA5 and NAM_ANL datasets. for cumulative area (ha−1 yr−1 ) Average exceedance area (km2 ) Cumulative exceedance in previous area (ha−1 yr−1 ) Average exceedance Average exceedance area (km2 ) in additional area (ha −1 yr−1 ) With LNG emissions Additional exceedance exceedance (ha−1 yr−1 ) Average Total area (km2 ) Without LNG NAM_ANL 5 1.0 0 - 1.2 5 1.2 ERA5 5 0.7 1 1.0 1.0 6 0.8 194 LNG SO2 emissions on the other hand, are expected to broaden aerial exceedance of critical load of acidity (CLA) by 3 km2 or roughly 30% increase (Fig. 8.6, Table 8.3). The baseline exceedance derives from emissions of a long-standing aluminum smelter nearby Canada LNG, and the additional area is mainly west of the Kitimat River (Fig. 8.6), again reflecting the deflection of air pollutant movement by steep topography. Nowhere else, including at the Bish Cove site of Kitimat LNG is S deposition exceedance expected. Indeed, modeled exceedance of CLA does not extend as far south as in Fig. 8.4, despite anticipated increased shipping activity and concomitant SO2 release. It is inferred that only areas in very close proximity of the existing smelter SO2 source will be subject to additional sulfur depsition from Canada LNG. The projection is that original CLA exceedances areas would be impacted by a further 4–6 kg ha−1 yr−1 of S due to LNG facilities. 8.4 Discussion Atmospheric chemistry modeling projected as much as 50% and 150 % increase in ambient SO2 and NO x respectively in the Terrace-Kitimat valley due to emissions from LNG projects. Along a north-south valley transect, these translate to maximum annual air concentrations of 5.2 ppb NOx and 1.2 ppb SO2 (Fig. 8.7) with LNG industry operations. Projected NOx concentrations are less than the monitored Canada-wide and BC averages of 7.8 and 9.2 ppb, respectively for NO2 , and much less than the 17 ppb national standard for 2020 (ECCC 2018b). The highest modeled SO2 concentration is slightly more than the 1.0 ppb Canada-wide and BC averages, but much less than the 5 ppb Canadian stan195 Figure 8.6 Spatial exceedance (modeled sulfur deposition minus 29 kg ha−1 yr−1 ) of CLA with, and without contributions from LNG industry using ERA5 and NAM_ANL datasets. dard (ECCC 2018b). The majority of modeled NOx and SO 2 concentrations, accounting for LNG industry emissions, are comparable to ambient levels that have been reported at locations 100-200 km from large industrial sources in Western Canada (e.g. at Fort Chipewyan in Alberta, NOx : 0.7–2.0 ppb, SO2 :< 0.8 ppb (Cho et al. 2017, Hsu 2013)). These suggest that LNG emissions in the immediate future would not result in violation of national and provincial objectives for atmospheric quality. Unlike many regions however, the study area is geographically-constrained and recent increased industrialization, 196 Table 8.3 Estimated exceedances of critical load (29 kg ha−1 yr−1 ) of sulfur deposition and without LNG emissions using ERA5 and NAM_ANL datasets. for cumulative area (ha−1 yr−1 ) Average exceedance area (km2 ) Cumulative exceedance in previous area (ha−1 yr−1 ) Average exceedance in additional area (ha −1 yr−1 ) area (km2 ) Average exceedance With LNG emissions Additional exceedance exceedance (ha−1 yr−1 ) Average Total area (km2 ) Without LNG NAM_ANL 10 33.5 3 5.0 37.1 13 29.7 ERA5 11 37.8 3 4.9 43.2 14 35.0 would require prolonged monitoring of pollutant concentrations for adaptive emissions management. Although the relative changes in ambient levels are greater than for SO2 , incremental NOx emissions were less significant of vegetation exposure. It is not likely that spatial exceedance of critical limits of NOx concentrations for sensitive lichen will occur with LNG operations, similar to projections by ESSA Technologies et al. (2014) and Williston et al. (2016). Virtually no additional area will exceed the critical load of N. Current and projected N exceedances in the limited Terrace area were fractions of the critical load which implied low increment in N accumulation. The assessment of little aerial impact of N deposition was due separate areas of the present, and future sources of NOx . Future LNG 197 Figure 8.7 Projected annual mean NOx and SO2 concentrations due to cumulative (total) emissions along a north-south transect through the TKV (right). facilities would be 60–75 km from the present urban core (Terrace) of vehicular NOx emissions. Industrial point emissions dilute and decrease their ambient concentration with longer travel distance. Consequently, the aggregate effect of NO x is less than it would be if emission stacks were closer to the principal area source. For SO2 on the other hand, new industrial source locations are nearer to an existing source of high emissions.The Canada LNG plant site for example, will be within 2 km of Rio Tinto’s smelter that is allowed up to 42 tonnes day −1 SO2 discharge, hence the modest extension of detrimental concentrations and deposition. Therefore, from an environmental sustainability perspective, clustering industrial activity vis-a-vis siting them at disparate, hitherto undisturbed 198 locations, should be carefully considered in planning and proposing new projects. Vegetation in an area of 20–22 km2 was projected to experience harmful exposure to total SO2 emissions. Routine inspection of foliar injury alongside tissue analysis of S content in hemlock needles in the valley, dating from several decades has continued to the present. The trend in recent years has been a rise in foliar sulfur with increasing SO2 emissions (Rio Tinto 2018), although the cause of higher occurrence of leaf chlorosis and necrosis on native plants in the vicinity of existing industrial source has often been unclear. However, on the evidence that modeled aerial exceedance is co-extensive with frequent identification of plant injury, direct SO2 exposure likely contributes to deleterious effects, and will be a major determinant of vegetation health in areas near proposed facilities. Further away (∼ 10 km) from project sites, there may be potential for enhanced sensitivity of conifers to pollution stress during needle elongation. Understory herbs are likely to be more impacted than trees especially in the growing season (late April to September) when leaf broadening and drier conditions facilitate leaf surface accumulation and intake. Adverse response would differ among floral components; however, except for the west-side-valley wall area around the shoreline, sub-alpine vegetation is particularly not threatened. The safety of high-elevation forest however, may not be absolute considering that exceedances were based on concentrations in ambient air (the lowest model layer) and not on in-cloud SO2 mixing ratios nor sulfate quantities in fog which are quite frequent in the Kitimat area. Due to aggregate industrial emissions, an area of 13–14 km2 was predicted to exceed the critical load of acidity for protection of forests on the predominantly mineral soils in the 199 study area. This estimate is within the range (1–28 km2 ) of areal exceedance projected by a government-led study with the CALPUFF model (ESSA Technologies et al. 2014). The CALPUFF study included emission scenarios from other projects (e.g. bitumen refinery, hydropower station) that are no longer seriously advanced by proponents. Apart from emission totals, differences in modeling technique and uncertainty in model outputs can lead to different projections. For instance, the CALPUFF study predicted concentrations that are approximately double (227 %) the measured SO2 concentrations. Further it assumed 80 ppb ozone concentration over the TKV, possibly resulting in high SO2 to sulfate conversion and total sulfur deposition rates than would occur when using site-specific ozone data. In the present study with CMAQ, gridded ozone mixing ratios are simulated in-situ, but hardly exceeded 40 ppb that approximates observed spring season maximum concentrations. CMAQ-predicted annual mean SO2 concentrations was roughly 50 % overestimation of Kitimat monitoring data. The indication thus, is that CMAQ acid deposition modeling is reasonably precise, whilst providing oxidant chemistry that is preferred (USEPA 2017) for regulatory assessment of multi-air pollutant discharges. How much restoration would be required for affected areas by acidification is as important as identifying terrestrial habitats at risk. Although cumulative areal CLA exceedance is small, projected area-weighted acidity loading of 29.7–35.0 kg S ha−1 yr−1 is slightly greater than the CLA itself (29.0 kg S ha−1 yr−1 ). This also, is about an order of magnitude greater than exceedances for forest soils reported for most parts of eastern North America (Duarte et al. 2013, McNulty et al. 2007), where acidifying industrial emissions has long been a major environmental challenge. Soils recovery from acidification has 200 been documented in regions that have experienced significant SO2 emissions reduction (Lawrence et al. 2015), but regeneration of original plant communities has also been less successful. Indeed, the danger of prolonged acidic deposition is the alteration of soil chemistry ultimately resulting to conditions that are favourable to fewer species. Because irreversible changes in habitat structure and composition can occur with new CLA exceedances, timely mitigation of acidic deposition is needed. Setting target loads of S and N deposition is one step to prevent adverse changes to the natural environment. Counteracting base cation depletion through liming could also be beneficial in maintaining the growth of native vegetation nearby LNG emission locations especially in view that acidifying intensity may span several future decades. With industrial development at pristine valley sites, it is important that the environment, including vegetation resources are fully profiled to ensure that future emissions are distant from the most valuable areas. 8.5 Concluding remarks Estimates of incremental impacts of major sources of atmospheric pollutants are useful, without which it will be difficult tracking the course of departures from environmental baselines. Emissions from two LNG projects in Kitimat will cause at least 50% and 150 % rise in ambient SO2 and NOx , respectively; however only an additional 4 km2 will be exposed to SO2 concentrations that are directly harmful to vegetation. An additional 3 km2 exceedance of critical load of acidity, with sulfur exceedance ranging 29.7–35.0 kg ha−1 yr−1 for a cumulative 13–14 km2 , suggested the importance of setting target loads and routine liming to prevent soil acidification and irreversible changes in vegetation compo201 sition in the vicinity of new emissions. Cumulative NO x concentrations are expected to remain below harmful levels; negligible (0–1 km2 ) increase in areal exceedance of nitrogen deposition is also anticipated. By defining aerial effects of acidifying LNG emissions on the environment ahead of actual releases, decisioning on liability for damage, compensation, restoration, etc. can be enriched. The environmental accountability perspective of mapping both incremental and cumulative impacts is therefore recommended for evaluations of future industrial emissions. 202 Chapter Nine Summary and Conclusion Optimization of a state-of-the-science platform required evaluating physics options and input data to extend the usefulness of recent information on critical levels and loads of atmospheric pollutants over an area with few ground measurements. The modeling for the industrializing Terrace-Kitimat valley, a physiographically complex region of northwestern British Columbia, Canada, was conducted at 1 km horizontal resolution with the WRF-SMOKE-CMAQ system. Investigations centered on representativeness of meteorological fields and air pollutant concentrations, evaluation of compliances to atmospheric quality standards, and quantification of acidification impact of industrial emissions. 9.0.1 Recapitulation of research findings A recap of findings of this research in the context of the questions in Chapter 1 is outlined: 1. What explains the behaviors of alternative PBL schemes in the WRF model for simulating the TKV’s surface meteorology? Analyses of spatial differences among PBL schemes for surface variables suggested some 203 factors that could have led to varying representations of meteorological variables that are of interest to atmospheric quality modeling. The coupling of PBL schemes to alternative surface-layer schemes which also differed in their diagnosis of variables in the lowest model layer such as skin temperature and friction velocity led to greater nighttime temperatures with the TKE-based schemes (MYNN3, MYJ and UW) than the first–order schemes. The inherent design for certain physical processes, for example moist convection and cloud-topped boundary layers with MYNN3 and UW PBL schemes was responsible for output of greater water vapor content with these schemes than those not specifically addressing these features e.g. the YSU and SH PBL schemes. Varying estimates of PBL depth attested to distinct methods for diagnosis of mixing heights, which also hinged on whether eddy transport across layers was modeled locally, nonlocally or in a hybrid fashion by individual schemes. The YSU and SH PBL schemes output similar properties apparently due to modeling at 1 km spatial resolution. 2. How does the choice of PBL scheme affect the simulation of PM2.5 , SO2 and NO2 concentrations and what reasons may account for differences in model outputs? For modeling the above air contaminants, it was found that the choice of PBL parameterization influenced quantitative and qualitative agreement with observational data. The top-ranked PBL schemes were MYNN3 for PM2.5 and MYJ for NO2 . Both schemes ranked high for absolute SO2 levels, but the MYJ and ACM2 schemes qualitatively better emulated peak summertime diurnal concentrations near elevated point sources in Kitimat. Greater nighttime SO2 concentrations with MYNN3 and the YSU PBL schemes, in less agreement with station monitoring 8 km downwind of emissions from tall stacks, sug204 gested sustained pollutant mixing and downward transport within the nocturnal boundary layer due to their representations of nonlocal mass flux transfers between model layers. With these schemes, inland penetrations of pollutant plumes were farther than those of ACM2, MYJ, and UW PBL schemes. For NO2 and PM2.5 that mainly discharged passively from fugitive, ground-level sources, hence are less accurately quantified than SO2 emissions, the MY- PBL schemes more reasonably reproduced peak season concentrations than other schemes. It was concluded that for air pollution modeling in rugged, remote areas, the mode of pollutant emissions is important for the choice of a PBL scheme. 3. Could quantile-based bias correction of CMAQ output improve usefulness for assessing regulatory compliance to air quality objectives? Linear regression formulas based on quantiles-mean matching of model output to observational data reduced quantitative biases in PM2.5 modeling especially for peak autumn and winter season concentrations. On average for year 2017, absolute errors in simulating annual means and 98th percentiles of daily averages, were 11 % and 10 %, respectively for bias-corrected outputs, compared to 45 % and 61 %, respectively for raw outputs. Initial non-correspondence in risk management categorization of pollutant exposure was corrected, thereby allowing for projections over unmonitored areas. 4. How well do simulations with various PBL parameterizations capture wet deposition of acidifying ions and what are the baseline exceedances of sulfur and nitrogen deposition in the TKV area? The skill of the WRF-SMOKE-CMAQ modeling system for wet deposition of acidifying 205 ions varied with nearness of measurement locations to major anthropogenic source of precursors and the quantities of individual species. Modeled quantities for sulfate and ammonium were more matched with observations at the Lakelse Lake station where the MYNN3, MYJ and YSU PBL schemes produced slight overestimations than at Haul Road station that is on the fence line of industrial pollutant sources. Normalized biases tended to be smaller for ammonium than for nitrate and sulfate that occurred in greater amounts. In 2017, forest soils were estimated to be in exceedance of critical load of acidity in the vicinity of the Rio Tinto smelter by 30.1–53.5 kg S ha −1 yr−1 . Exceedance of critical load of nutrient nitrogen restricted to the Terrace area (≤ 7 km2 ) ranged between none and 0.71 kg ha −1 yr−1 . 5. What choices of atmospheric forcing data and PBL scheme in the WRF model can reproduce precipitation for the purpose of projecting atmospheric deposition over the TKV and what is the uncertainty in the precipitation field with the best fit simulations? The MYJ and MYNN3 PBL schemes that best quantified ambient concentrations of air pollutants in the valley alongside three coarse-resolution atmospheric datasets that are available in formats compatible with the WRF model were further assessed for how closely their outputs emulate spatial and temporal distribution of precipitation. Underestimation by more than 40 % on average of total precipitation ranging 1170–2380 mm in 2017, was found for simulations using either NARR for atmospheric forcing or MYNN3 as a PBL scheme. The large precipitation undercapture with this combination likely stemmed from little assimilation of observational data in the reanalyses and poor adaptation of the MYNN3 PBL scheme for frontal precipitation that dominate the wet months. With 206 NAM_ANL and ERA5 as atmospheric forcing data and MYJ as the PBL scheme, the uncertainty in annual simulated precipitation amount ranged between 40 % overestimation and 20 % underestimation of observational data. 6. What incremental and aggregate impacts could arise should the existing smelter at Kitimat emit 42 tonnes day−1 that is the highest permissible rate? An additional 15 tonnes daily SO2 emission was projected to result in new aerial exceedances of 6– 7 km2 and 3–4 km2 of the critical level of plant exposure and critical load of acidity respectively. Exceedances of SO2 concentrations and sulfur deposition over pre-existing impacted areas were predicted to rise by 21–27 % and 30–33 %, respectively. Cumulatively, 16–18 km2 of plant habitat and 10–11 km2 of soil in an area contiguous with the smelter site will likely be damaged by its SO2 emission under the latest regulation. 7. For proposed developments in the TKV, (a) To what additional extent could vegetation be exposed to harmful NOx and SO2 concentrations from LNG projects? (b) How much direct vegetation exposure to harmful NOx and SO2 concentrations could result from aggregate industrial emissions? (c) By how much could soil nitrogen enrichment and acidification change as a result of the LNG industry? (d) What aerial exceedances of critical loads of nitrogen and sulfur deposition will arise from aggregate industrial emissions? 207 In the event that two LNG projects are built and in operation, an area of 4 km2 could further be exposed to SO2 concentrations that are directly harmful to vegetation, with an average exceedance of 0.7–0.8 ppb in the new areas and 17–27 % increased magnitude over pre-existing impacted areas. For aggregate emissions, a 20-22 km2 area above SO2 critical limit for lichen, with average exceedance of 3.5–3.7 ppb was predicted. Cumulative NO x concentrations are expected to remain below harmful levels. For indirect vegetation impacts via soil pollution, pre-existing areal exceedance of nitrogen will barely increase (0–1 km2 ), with overall average exceedance of 0.8–1.2 kg ha−1 yr−1 . Additional 3 km2 new aerial exceedance of critical load of acidity, for a cumulative 13–14 km2 area and 29.7–35.0 kg ha−1 yr−1 average exceedance was also predicted. These projections assume that new LNG facilities will emit all NOx , SO2 and other air pollutants from elevated point sources. 9.0.2 Significance of study findings By comparing model outputs with actual measurements, this study guides what configurations in the WRF meteorological driver are more appropriate for the northwest BC coast that is dominated by sharp elevation gradients, and frequently humid conditions. This importance is in view of errors that could arise from inexactness of emissions input, terrain representation, and chemical reaction chains that are processed within CMAQ. Especially for the major air pollutants (SO2 , NO2 , PM 2.5 ), the evaluations of whether the CMAQ model captures observed quantities for various PBL parameterizations are useful. PBL schemes development and verification have mainly been for numerical weather prediction; their realisms of chemical transport and dispersion seldom investigated. The 208 analyses of model outputs for separate runs and different peak seasons provide ample information on their dependencies on the weather, type of pollutant source, and topography. For example, it was found for the various air pollutants that the UW PBL parameterization was most different from other PBL schemes due to these factors. This work contributes significantly to atmospheric science literature since to the author’s knowledge, no prior modeling study of PBL schemes influence on air quality variables over fjord valleys exists. Model evaluations in this study can be helpful to CMAQ users in general, although the use of a simple, comprehensible, post-processing technique to ameliorate the initial bias in model output as was done for PM2.5 , has particular relevance for airshed managers. In objectively adjusting modeled concentrations, with corrections spread to all grid cells, a realistic classification of pollutant exposure was achieved. That demonstration, it is believed, will benefit compliance assessment of pollutant concentrations especially if relying on the output from a single CMAQ configuration. Projected ambient pollutant concentrations and acidifying deposition from this study may provide a reference for changes in the atmospheric quality of similarly isolated areas that are planned for industrialization. It however contributes lastingly through mapping pre-existing levels of major pollutants in the TKV— an exercise that was not realized until this study. The TKV, like much of northern BC is sparsely monitored for air quality. However, retrospective modeling with CMAQ allowed not only the quantification of preexisting pollutant levels from natural and anthropogenic sources, but also the estimation of spatial exceedances of safe limits for the forest ecosytem. Future monitoring and mod209 eling data may be compared to them for changes in the spatial distribution of pollutants. In this regard, the study’s findings are useful for effective planning and management of air emissions in the TKV. 9.0.3 Directions for future research Pollutant concentrations and deposition estimates were from the best fit simulations. Appropriate choice of model physics, however, does not preclude the need for continuous improvement of initial and boundary source fields. Availability of atmospheric reanalyses at 1 km horizontal gridding for instance would apart from obviating downscaling errors and enhancing the precision of numerical schemes, reduce computational costs. Similar improvement for area source emissions inventorying would also facilitate spatiotemporal gridding with the SMOKE tool. With more refining and updating of atmospheric modeling systems, a top priority would be their validation in areas with geographical similarity to the TKV, to allow for more comparisons and creation of unified benchmarks for high-resolution air quality modeling over complex orography. High-resolution CMAQ simulations are fundamental to precise projections of future scenarios and the approach of mapping atmospheric deposition exceedances, with and without emission increase is applicable to areas where the permitting of industrial projects are phased. Even for the TKV area itself, chemical transport models can also linked to other models to investigate emissions effects on aquatic environment, public health, atmospheric visibility, etc. It is necessary that predictive modeling of big industrial projects utilize updated science tools to enrich information about the extent and magnitude of 210 environmental impacts. 211 References Aguilera, R. F., J. Inchauspe, and R. D. Ripple, 2014: The Asia Pacific natural gas market: Large enough for all? Energy Policy, 65, 1–6, doi:10.1016/ j.enpol.2013.10.014. Andrews, A., and R. K. Lattanzio, 2013: Petroleum coke: industry and environmental issues. R43263, http://petcoke.dev.psiclientweb.com/wp-content/uploads/ 2014/06/nps71-120213-07.pdf. Appel, K. W., S. J. Roselle, R. C. Gilliam, and J. E. Pleim, 2010: Sensitivity of the community multiscale air quality (CMAQ) model v4.7 results for the eastern United States to MM5 and WRF meteorological drivers. Geosci. Model Dev., 3 (1), 169–188, doi:10.5194/gmd-3-169-2010. Appel, K. W., and Coauthors, 2018: Overview and evaluation of the Community Multiscale Air Quality (CMAQ) modeling system version 5.2. 69–73, doi:10.1007/ 978-3-319-57645-9_11. Arakawa, A., J.-H. Jung, and C.-M. Wu, 2011: Toward unification of the multiscale 212 modeling of the atmosphere. Atmos. Chem. Phys, 11, 3731–3742, doi:10.5194/acp-113731-2011. Atkinson, R. W., I. C. Mills, H. A. Walton, and H. R. Anderson, 2015: Fine particle components and health—a systematic review and meta-analysis of epidemiological time series studies of daily mortality and hospital admissions. J. Expo. Sci. Environ. Epidemiol., 25 (2), 208–214, doi:10.1038/jes.2014.63. Augustaitis, A., 2011: Impact of meteorological parameters on responses of pine crown condition to acid deposition at aukštaitija national park. Balt. For., (17), 205–214, https://www.balticforestry.mi.lt/bf/PDF_Articles/2011- 17[2 ]/Augustaitis_201117(2)_205_214.pdf. Avolio, E., S. Federico, M. Miglietta, T. L. Feudo, C. Calidonna, and A. Sempreviva, 2017: Sensitivity analysis of WRF model PBL schemes in simulating boundary-layer variables in southern Italy: An experimental campaign. Atmos. Res., 192, 58–71, doi: 10.1016/j.atmosres.2017.04.003. Balzarini, A., and Coauthors, 2014: Sensitivity analysis of PBL schemes by comparing WRF model and experimental data. Geosci. Model Dev. Discuss., 7, 6133–6171, doi:10.5194/gmdd-7-6133-2014. Banks, R. F., and J. M. Baldasano, 2016: Impact of WRF model PBL schemes on air quality simulations over Catalonia, Spain. Sci. Total Environ., 572, 98–113, doi: 10.1016/j.scitotenv.2016.07.167. BC Oil and Gas Commission, 2020: 2017 oil and gas reserves and production report. 213 BCMOECS, 2020: Provincial air quality objective for pm2.5 . https: //www2.gov.bc.ca/gov/content/environment/air-land-water/air/air-qualitymanagement/regulatory-framework/objectives-standards/pm2-5. Beck, H. E., and Coauthors, 2017: Global-scale evaluation of 22 precipitation datasets using gauge observations and hydrological modeling. Hydrol. Earth Syst. Sci., 21 (12), 6201–6217, doi:10.5194/hess-21-6201-2017. Belo-Pereira, M., E. Dutra, and P. Viterbo, 2011: Evaluation of global precipitation data sets over the Iberian peninsula. J. Geophys. Res., 116, D20 101, doi: 10.1029/2010jd015481. Besse, W. J., 2010: Old-growth specklebelly lichen surveys – fia final report 2009/10. 41 pp, https://www.for.gov.bc.ca/hfd/library/FIA/2010/LBIP_6906008.pdf. Blande, J. D., J. K. Holopainen, and U. Niinemets, 2014: Plant volatiles in polluted atmospheres: stress responses and signal degradation. Plant Cell Environ., 37 (8), 1892–1904, doi:10.1111/pce.12352. Bobbink, R., and Coauthors, 2010: Global assessment of nitrogen deposition effects on terrestrial plant diversity: a synthesis. Ecol. Appl., (20), 30–59, doi:10.1890/081140.1, https://doi.org/10.1890/08-1140.1. Boylan, J. W., and A. G. Russell, 2006: PM and light extinction model performance metrics, goals, and criteria for three-dimensional air quality models. Atmos. Environ., 40 (26), 4946–4959, doi:10.1016/j.atmosenv.2005.09.087. 214 BP, 2020: Energy outlook: demand by fuel. https://www.bp.com/en/global/ corporate/energy-economics/energy-outlook/demand-by-region.html. Bretherton, C. S., and S. Park, 2009: A new moist turbulence parameterization in the community atmosphere model. J. Climate, 22, 3422–3448, doi:10.1175/ 2008JCLI2556.1. Brook, J., S. Cober, M. Freemark, T. Harner, S. Li, J. Liggio, P. Makar, and B. Pauli, 2019: Advances in science and applications of air pollution monitoring: A case study on oil sands monitoring targeting ecosystem protection. J. Air Waste Manage. Assoc., 69 (6), 661–709, doi:10.1080/10962247.2019.1607689. Camuffo, D., 2014: Atmospheric stability and pollutant dispersion. In: Microclimate for Cultural Heritage. 245–282, doi:10.1016/B978-0-444-63296-8.00008-1. Carolina Environmental Program, 2009: SMOKE v2.6 user’s manual. The University of North Carolina at Chapel Hill, https://www.cmascenter.org/smoke/ documentation/2.6/html/. CCME, 2012: Guidance document on air zone management. https:// www.ccme.ca/files/Resources/air/aqms/pn_1481_gdazm_e.pdf. CCME, 2014a: 2008-09 Acid Rain Strategy Progress Report. https://www.ccme.ca/ files/Resources/air/acid_rain/2008_09_ar_progress_rpt_pn1458_e.pdf. CCME, 2014b: AQMS. http://www.ccme.ca/en/resources/air/aqms.html. 215 Center for Liquefied Natural Gas, 2020: What is lng? https://lngfacts.org/aboutlng/what-is-lng/. Chang, J. C., and S. R. Hanna, 2004: Air quality model performance evaluation. Meteorol. Atmos. Phys., 87 (1–3), 167–196, doi:10.1007/s00703-003-0070-7. Cheng, F.-Y., S.-H. Chin, and T.-H. Liu, 2012: The role of boundary layer schemes in meteorological and air quality simulations of the taiwan area. Atmos. Environ., 54, 714–727, doi:10.1016/j.atmosenv.2012.01.029. Cheng, L., H. Sandhu, R. Angle, K. McDonald, and R. Myrick, 2000: Rural particulate matter in alberta, canada. Atmos. Environ., 34 (20), 3365–3372, doi:10.1016/ s1352-2310(00 )00123-0. Cho, S., K. Vijayaraghavan, D. Spink, J. Jung, R. Morris, and R. Pauls, 2017: Assessment of regional acidifying pollutants in the Athabasca Oil Sands area under different emission scenarios. Atmos. Environ., 156, 160–168, doi:10.1016/ j.atmosenv.2017.02.038. Chumanová-Vávrová, E., O. Cudlín, and P. Cudlín, 2015: Spatial and temporal patterns of ground vegetation dominants in mountain spruce forests damaged by sulphur air pollution (Giant mountains, Czech Republic). Boreal Environ. Res., 20, 620–636, https://helda.helsinki.fi/bitstream/10138/228299/1/ber20-5-620.pdf. Cirtina, D., N. Ionescu, and L. Cirtina, 2016: Environmental impact as- sessment related to metallurgical industry activities. Metalurgija, 55, 482–484, https://hrcak.srce.hr/153690. 216 Coats Jr., C. L., 2017: The BAMS/EDSS/Models-3 I/O API: User Manual. https: //www.cmascenter.org/ioapi/documentation/all_versions/html/AA.html. Cohen, A., S. M. Cavallo, M. C. Coniglio, and H. E. Brooks, 2015: A review of planetary boundary layer parameterization schemes and their sensitivity in simulating southeastern U.S. cold season severe weather environments. Wea. Forecasting, 30, 591–612, doi:10.1175/WAF-D-14-00105.1. Committee on the Status of Endangered Wildlife in Canada, 2006: Assessment and update status report on the cryptic paw Nephroma occultum in Canada. 28 pp, https://www.sararegistry.gc.ca/virtual_sara/files/cosewic/sr_cryptic_paw_ lichen_e.pdf. Coordination Centre for Effects, 2017: Manual for modelling and mapping critical loads & levels. Chapters 3 and 5, https://www.umweltbundesamt.de/en/ manual-for-modelling-mapping-critical-loads-levels. Cosby, B. J., G. M. Hornberger, J. N. Galloway, and R. E. Wright, 1985: Time scales of catchment acidification. a quantitative model for estimating freshwater acidification. Environ. Sci. Technol., 19 (12), 1144–1149, doi:10.1021/es00142a001. Crouse, D. L., and Coauthors, 2012: Risk of nonaccidental and cardiovascular mortality in relation to long-term exposure to low concentrations of fine particulate matter: A canadian national-level cohort study. Environ. Health Perspect., 120 (5), 708–714, doi:10.1289/ehp.1104049. 217 Dennis, R., and Coauthors, 2010: A framework for evaluating regional scale numerical photochemical modelling systems. Environ. Fluid Mech., 10, 471–489, doi: 10.1007/s10652-009-9163-2. Dietz, T., 2003: The struggle to govern the commons. Science, 302 (5652), 1907–1912, doi:10.1126/science.1091015. Dise, N. B., and Coauthors, 2011: Nitrogen as a threat to european terrestrial biodiversity. The European Nitrogen Assessment, M. A. Sutton, C. M. Howard, J. W. Erisman, G. Billen, A. Bleeker, P. Grennfelt, H. van Grinsven, and B. Grizzetti, Eds., Cambridge University Press, 463–494, doi:10.1017/cbo9780511976988.023. District of Kitimat, 2020: Current major projects. https://www.kitimat.ca/en/ business-and-development/current-major-projects.aspx. Djalalova, I., L. D. Monache, and J. Wilczak, 2015: PM2.5 analog forecast and kalman filter post-processing for the community multiscale air quality (CMAQ) model. Atmos. Environ., 108, 76–87, doi:10.1016/j.atmosenv.2015.02.021. Driscoll, C. T., and Coauthors, 2001: Acidic deposition in the northeastern United States: Sources and inputs, ecosystem effects, and management strategies. BioScience, 51 (3), 180–198, doi:10.1641/0006-3568(2001)051[0180:aditnu]2.0.co;2. Du, E., M. E. Fenn, W. D. Vries, and Y. S. Ok, 2019: Atmospheric nitrogen deposition to global forests: Status, impacts and management options. Environ. Pollut., 250, 1044–1048, doi:10.1016/j.envpol.2019.04.014. 218 Duan, L., Q. Yu, Q. Zhang, Z. Wang, Y. Pan, T. Larssen, J. Tang, and J. Mulder, 2016: Acid deposition in Asia: Emissions, deposition, and ecosystem effects. Atmos. Environ., 146, 55–69, doi:10.1016/j.atmosenv.2016.07.018. Duarte, N., L. H. Pardo, and M. J. Robin-Abbott, 2013: Susceptibility of forests in the northeastern USA to nitrogen and sulfur deposition: Critical load exceedance and forest health. Water Air Soil Pollut., 224 (2), 1355, doi:10.1007/s11270-012-1355-6. ECCC, 1991: The potential of soils and bedrock to reduce the acidity of atmospheric deposition. http://ftp.geogratis.gc.ca/pub/nrcan_rncan/raster/atlas_5_ed/eng/ environment/ecology/mcr4157.jpg. ECCC, 2018a: port 2016). Canada–United States Air Quality Agreement (Progress re- https://www.canada.ca/en/environment-climate-change/services/ air-pollution/publications/canada-united-states-air-quality-report-2016.html. ECCC, 2018b: Canadian environmental sustainability indicators: Air qual- ity. https://www.canada.ca/content/dam/eccc/documents/pdf/cesindicators/ air-quality/air-quality-en.pdf. ECCC, 2019: Canadian climate normals 1981-2010 station data. http: //climate.weather.gc.ca/climate_normals/results_1981_2010_e.html?stnID= 402&autofwd=1. ECCC, 2020: Air Pollutants Emissions Inventory online search. https://pollutionwaste.canada.ca/air-emission-inventory/. 219 ECMWF, 2019: ERA5 data documentation. https://confluence.ecmwf.int/ display/CKB/ERA5datadocumentation. Emery, C., Z. Liu, A. G. Russell, M. T. Odman, G. Yarwood, and N. Kumar, 2017: Recommendations on statistics and benchmarks to assess photochemical model performance. J. Air Waste Manage. Assoc., 67 (5), 582–598, doi:10.1080/ 10962247.2016.1265027. Emery, C. A., E. Tai, and G. Yarwood, 2001: Enhanced meteorologi- cal modeling and performance evaluation for two Texas ozone episodes. https://www.tceq.texas.gov/assets/public/implementation/air/am/contracts/ reports/mm/EnhancedMetModelingAndPerformanceEvaluation.pdf. Environmental Appeal Board, 2015: Decisions: Emily Toews; Elisabeth Stannus v. Director, Environmental Management Act (Rio Tinto Alcan Inc., Third Party). 1–2, 58–84 pp, http://www.eab.gov.bc.ca/ema/2013ema007g_010g_ corrigendum_2.pdf. Environmental Modeling Center, 2017: Mesoscale model/analysis systems faq. https://www.emc.ncep.noaa.gov/mmb/research/FAQ-meso.html. ESSA Technologies, J. Laurence, Limnotek, Risk Sciences International, Trent University, and Trinity Consultants, 2014: Kitimat airshed emissions effects assessment. 205 pp, https://www2.gov.bc.ca/assets/gov/environment/air-land-water/air/ reports-pub/2014-04-25_essa-qa_kitimat_airshed.pdf. 220 ESSA Technologies, J. Laurence, Risk Sciences International, Trent University, and Trinity Consultants, 2016: Supplementary report to the Prince Rupert Airshed study. 98 pp, https://www2.gov.bc.ca/assets/gov/environment/air-land-water/ air/reports-pub/pr-airshed-study-report-summ.pdf. ESSA Technologies, Risk Sciences International, and Trinity Consultants, 2015: Pm2.5 supplement to the Kitimat Airshed emissions effects assessment. https://www2.gov.bc.ca/assets/gov/environment/air-land-water/air/reportspub/2015-03-31_essa-pm2-5_suppl-kaeea.pdf, 19 pp. Evans, J. P., M. Ekström, and F. Ji, 2011: Evaluating the performance of a WRF physics ensemble over south-east Australia. Climate Dyn., 39 (6), 1241–1258, doi: 10.1007/s00382-011-1244-5. Fernández-González, S., F. Valero, J. L. Sánchez, E. Gascón, L. López, E. GarcíaOrtega, and A. Merino, 2015: Numerical simulations of snowfall events: Sensitivity analysis of physical parameterizations. J. Geophys. Res. Atmos., 120, 10 130–10 148, doi:10.1002/2015jd023793. Food and Agricultural Organization, 1989: Arid zone forestry - A guide for field technicians. Table 4.1, http://www.fao.org/3/T0122E/t0122e08.htm. García-Díez, M., J. Fernández, L. Fita, and C. Yagüe, 2013: Seasonal dependence of WRF model biases and sensitivity to PBL schemes over. Europe. Quart. J. Roy. Meteor. Soc., 139, 501–514, doi:10.1002/qj.1976. 221 Gerboles, M., and H. I. Reuter, 2010: Estimation of the measurement uncertainty of ambient air pollution datasets using geostatistical analysis. 38 pp, http:// publications.jrc.ec.europa.eu/repository/bitstream/JRC59441/lbna24475enc.pdf. Gheorge, I. F., and B. Ion, 2011: The effects of air pollutants on vegetation and the role of vegetation in reducing atmospheric pollution. In: The Impact of Air Pollution on Health, Economy, Environment and Agricultural sources. 241–280, doi: 10.5772/17660. Gilliam, F., 2019: Excess nitrogen in temperate forest ecosystems decreases herbaceous layer diversity and shifts control from soil to canopy structure. Forests, 10, 66, doi:10.3390/f10010066. Grell, G. A., and S. R. Freitas, 2014: A scale and aerosol aware stochastic convective parameterization for weather and air quality modeling. Atmos. Chem. Phys., 14, 5233–5250, doi:10.5194/acp-14-5233-2014. Gunwani, P., and M. Mohan, 2017: Sensitivity of WRF model estimates to various PBL parameterizations in different climatic zones over India. Atmos. Res., 194, 43–65, doi:10.1016/j.atmosres.2017.04.026. Habashi, F., 2011: Pollution problems in the metallurgical industry: A review. J. Mining Environ., 2 (1), 17–26, doi:10.22044/jme.2012.16. Hariprasad, K. B. R. R., C. V. Srinivas, A. B. Singh, S. V. B. Rao, R. Baskaran, and B. Venkatraman, 2014: Numerical simulation and intercomparison of boundary 222 layer structure with different PBL schemes in WRF using experimental observations at a tropical site. Atmos. Res., 145, 27–44, doi:10.1016/j.atmosres.2014.03.023. Hedley, M., and D. L. Singleton, 1997: Evaluation of an air quality simulation of the lower fraser valley-i. meteorology. Atmos. Environ., 31, 1605–1615, doi:10.1016/ S1352-2310(96. Hong, S.-Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an explicit treatment of entrainment processes. Mon. Wea. Rev., 134, 2318–2341, doi: 10.1175/MWR3199.1. Hsu, Y.-M., 2013: Trends in passively-measured ozone, nitrogen dioxide and sulfur dioxide concentrations in the Athabasca oil sands region of Alberta, Canada. Aerosol Air Qual. Res., 13 (5), 1448–1463, doi:10.4209/aaqr.2012.08.0224. Hu, J., J. Chen, Q. Ying, and H. Zhang, 2016: One-year simulation of ozone and particulate matter in China using WRF/CMAQ modeling system. Atmos. Chem. Phys., 16 (16), 10 333–10 350, doi:10.5194/acp-16-10333-2016. Hu, X.-M., J. Nielsen-Gammon, and F. Zhang, 2010: Evaluation of three planetary boundary layer schemes in the WRF model. J. Appl. Meteor. Climatol., 49, 1831–1843, doi:10.1175/2010JAMC2432.1. Hu, X.-M., M. Xue, R. A. McPherson, E. Martin, D. H. Rosendahl, and L. Qiao, 2018: Precipitation dynamical downscaling over the Great Plains. J. Adv. Model. Earth Syst., 10 (2), 421–447, doi:10.1002/2017ms001154. 223 Hu, Y., N. Bellaloui, M. Tigabu, J. Wang, J. Diao, K. Wang, R. Yang, and G. Sun, 2015: Gaseous NO2 effects on stomatal behavior, photosynthesis and respiration of hybrid poplar leaves. Acta Physiol. Plant, 37, 39, doi:10.1007/s11738-014-1749-8, https://doi.org/10.1007/s11738-014-1749-8. Iacono, M. J., J. S. Delamere, E. J. Mlawer, M. W. Shephard, S. A. Clough, and W. D. Collins, 2008: Radiative forcing by long-lived greenhouse gases: Calculations with the aer radiative transfer models. J. Geophys. Res., 113, D13 103, doi: 10.1029/2008JD009944 |. Impact Assessment Agency of Canada, 2019: Kitimat LNG expansion poject description summary. https://iaac-aeic.gc.ca/050/documents/p80180/130805E.pdf. Janjić, Z. I., 1994: The step-mountain Eta coordinate model: Further developments of the convection, viscous sublayer, and turbulence closure schemes. Mon. Wea. Rev., 122, 927–945, doi:10.1175/1520-0493. Janjić, Z. I., 2002: Nonsingular implementation of the Mellor-Yamada level 2.5 scheme in the NCEP meso model. 437, NCEP Office Note, 61 pp, https:// repository.library.noaa.gov/view/noaa/11409. Jiménez, P. A., J. Dudhia, J. F. Gonzalez-Rouco, J. Navarro, J. P. Montavez, and E. García-Bustamante, 2012: A revised scheme for the WRF surface layer formulation. Mon. Wea. Rev., 140, 898–918, doi:10.1175/MWR-D-11-00056.1. Josipovic, M., H. J. Annegarn, M. A. Kneen, J. J. Pienaar, and S. J. Piketh, 2011: Atmospheric dry and wet deposition of sulphur and nitrogen species and assess224 ment of critical loads of acidic deposition exceedance in South Africa. S. Afr. J. Sci., 107 (3/4), ID 478, doi:10.4102/sajs.v107i3/4.478. Kelly, J. T., and Coauthors, 2018: Modeling NH4 NO3 over the San Joaquin Valley during the 2013 DISCOVER-AQ campaign. J. Geophys. Res. Atmos., 123 (9), 4727– 4745, doi:10.1029/2018jd028290. Klock, R., and J. Mullock, 2000: The Weather of British Columbia: Graphic Area Forecast 31. 61–73 pp., http://www.navcanada.ca/EN/media/Publications/ LocalAreaWeatherManuals/LAWM-BC-EN.pdf. Kota, S. H., and Coauthors, 2018: Year-long simulation of gaseous and particulate air pollutants in India. Atmos. Environ., 180, 244–255, doi:10.1016/ j.atmosenv.2018.03.003. Krzyzanowski, J., 2010: Approaching cumulative effects through air pollution modelling. Water Air Soil Pollut., 214 (1-4), 253–273, doi:10.1007/s11270-010-0421-1. Kurt, O. K., J. Zhang, and K. E. Pinkerton, 2016: Pulmonary health effects of air pollution. Curr. Opin. Pulm. Med., 22, 138–43, doi:10.1097/MCP.0000000000000248. Lange, O. S., 2003: Regional weather guides: northern mainland coast. In: Living with weather along the British Columbia Coast: the veil of chaos. Environment Canada, 174–187 pp. Lawrence, G. B., P. W. Hazlett, I. J. Fernandez, R. Ouimet, S. W. Bailey, W. C. Shortle, K. T. Smith, and M. R. Antidormi, 2015: Declining acidic deposition begins reversal 225 of forest-soil acidification in the northeastern U.S. and eastern Canada. Environ. Sci. Technol., 49 (22), 13 103–13 111, doi:10.1021/acs.est.5b02904. Lee, J., H. H. Shin, S.-Y. Hong, P. A. Jiménez, J. Dudhia, and J. Hong, 2015: Impacts of subgrid-scale orography parameterization on simulated surface layer wind and monsoonal precipitation in the high-resolution WRF model. J. Geophys. Res. Atmos., 120 (2), 644–653, doi:10.1002/2014jd022747. Legge, A. H., and S. V. Krupa, 2002: Effects of sulphur dioxide. In: Air Pollution and Plant Life (second edition). Wiley, Chichester, 135–162 pp. Levasseur, P. A., S. A. Watmough, J. Aherne, C. J. Whitfield, and M. C. Eimers, 2020: Estimating mineral surface area and base cation weathering rates of spodosols under forest in British Columbia, Canada. Geoderma Reg., 20, e00 247, doi:10.1016/ j.geodrs.2019.e00247. Li, L., W. Li, and J. Jin, 2014: Improvements in WRF simulation skills of southeastern United States summer rainfall: Physical parameterization and horizontal resolution. Climate Dyn., 43 (7-8), 2077–2091, doi:10.1007/s00382-013-2031-2. Li, P., 2005: Box-cox transformations: An overview. https://www.ime.usp.br/ ~abe/lista/pdfm9cJKUmFZp.pdf. Li, T., H. Wang, T. Zhao, M. Xue, Y. Wang, H. Che, and C. Jiang, 2016: The impacts of different PBL schemes on the simulation of PM2.5 during severe haze episodes in the Jing-Jin-Ji region and its surroundings in China. Adv. in Meteorol., ID 6295 878, 15 pp, doi:10.1155/2016/6295878. 226 Li, Y., M. C. Barth, and A. L. Steiner, 2019: Comparing turbulent mixing of atmospheric oxidants across model scales. Atmos. Environ., 199, 88–101, doi:10.1016/ j.atmosenv.2018.11.004. Lilhare, R., S. J. Déry, S. Pokorny, T. A. Stadnyk, and K. A. Koenig, 2019: Intercomparison of multiple hydroclimatic datasets across the Lower Nelson River Basin, Manitoba, Canada. Atmos. Ocean, 57 (4), 262–278, doi:10.1080/ 07055900.2019.1638226. Lin, Y., and K. E. Mitchell, 2005: The NCEP stage II/IV hourly precipitation analyses: Development and applications. http://citeseerx.ist.psu.edu/viewdoc/ download?doi=10.1.1.182.2080&rep =rep1&type=pdf. Liu, J., M. Bray, and D. Han, 2012: Sensitivity of the weather research and forecasting (WRF) model to downscaling ratios and storm types in rainfall simulation. Hydrol. Process., 26 (20), 3012–3031, doi:10.1002/hyp.8247. Longley, H., 2019: Conflicting interests: Development politics and the environmental regulation of the Alberta oil sands industry, 1970–1980. Environ. Hist. Camb., doi: 10.3197/096734019x15463432086919. Lyu, B., Y. Zhang, and Y. Hu, 2017: Improving PM2.5 air quality model forecasts in china using a bias-correction framework. Atmosphere, 8 (12), 147, doi:10.3390/ atmos8080147. Makar, P. A., and Coauthors, 2018: Estimates of exceedances of critical loads for 227 acidifying deposition in Alberta and Saskatchewan. Atmos. Chem. Phys., 18 (13), 9897–9927, doi:10.5194/acp-18-9897-2018. McKendry, I. G., 2006: Background concentrations of PM2.5 and ozone in British Columbia, Canada. http://citeseerx.ist.psu.edu/viewdoc/download?doi= 10.1.1.510.4465&rep =rep1&type=pdf. McNulty, S. G., E. C. Cohen, J. A. M. Myers, T. J. Sullivan, and H. Li, 2007: Estimates of critical acid loads and exceedances for forest soils across the conterminous United States. Environ. Pollut., 149 (3), 281–292, doi:10.1016/j.envpol.2007.05.025. Meidinger, D., 2018: Terrestrial Vegetation of British Columbia. Lab for Advanced Spatial Analysis, Department of Geography, University of British Columbia, Vancouver, https://ibis.geog.ubc.ca/biodiversity/eflora/TerrestrialVegetation.html. Mesinger, F., and Coauthors, 2006: North American Regional Reanalysis. Bull. Amer. Meteor. Soc., 87 (3), 343–360, doi:10.1175/bams-87-3-343. Miao, J.-F., and Coauthors, 2007: Evaluation of MM5 mesoscale model at local scale for air quality applications over the Swedish west coast: Influence of PBL and LSM parameterizations. Meteorol. Atmos. Phys., 99 (1-2), 77–103, doi:10.1007/s00703-0070267-2. Misenis, C., and Y. Zhang, 2010: An examination of sensitivity of WRF/Chem predictions to physical parameterizations, horizontal grid spacing, and nesting options. Atmos. Res., 97, 315–334, doi:10.1016/j.atmosres.2010.04.005. 228 Mohan, M., and M. Gupta, 2018: Sensitivity of PBL parameterizations on PM1 0 and ozone simulation using chemical transport model WRF-Chem over a sub-tropical urban airshed in India. Atmos. Environ., 185, 53–63, doi:10.1016/ j.atmosenv.2018.04.054. Moran, M. D., and Coauthors, 2008: Predicted acid deposition critical-load exceedances across canada from a one-year simulation with a regional particulatematter model. 20 pp, http://ams.confex.com/ams/pdfpapers/132916.pdf. Moya-Álvarez, A. S., R. Estevan, S. Kumar, J. L. F. Rojas, J. J. Ticse, D. MartínezCastro, and Y. Silva, 2020: Influence of PBL parameterization schemes in WRF_ARW model on short-range precipitation's forecasts in the complex orography of Peruvian Central Andes. Atmos. Res., 233, 104 708, doi:10.1016/ j.atmosres.2019.104708. Mues, A., A. Laurer, A. Lupascu, M. Rupakheti, F. Kuik, and M. G. Lawrence, 2018: WRF and WRF-Chem v3.5.1 simulations of meteorology and black carbon concentrations in the Kathmandu Valley. Geosci. Model Dev, 11, 2067–2091, doi: 10.5194/gmd-11-2067-2018. Nakanishi, M., and H. Niino, 2009: Development of an improved turbulence closure model for the atmospheric boundary layer. J. Meteor. Soc. Japan, 87, 895–912, doi:10.2151/jmsj.87.895. Nasr, M., M. Castonguay, J. Ogilvie, B. A. Raymond, and P. A. Arp, 2010: Modelling and mapping critical loads and exceedances for the georgia basin, british columbia 229 using a zero base-cation depletion criterion. J. Limnol., 69, 181–192, doi:10.4081/ jlimnol.2010.s1.181. Neal, L., P. Agnew, S. Moseley, C. Ordóñez, N. Savage, and M. Tilbee, 2014: Application of a statistical post-processing technique to a gridded, operational, air quality forecast. Atmos. Environ., 98, 385–393, doi:10.1016/j.atmosenv.2014.09.004. Nielsen-Gammon, J. W., X. M. Hu, F. Q. Zhang, and J. E. Pleim, 2010: Evaluation of planetary boundary layer scheme sensitivities for the purpose of parameter estimation. Mon. Weather Rev., 138, 3400–3417, doi:10.1175/2010MWR3292.1. Olson, J. B., J. S. Kenyon, W. M. Angevine, J. M. Brown, M. Pagowski, and K. Sušelj, 2019: A description of the Mynn-Edmf scheme and the coupling to other components in WRF-ARW. NOAA Technical Memorandum OAR GSD, 61, 37 pp, doi: 10.25923/n9wm-be49. Otte, T. L., and J. E. Pleim, 2010: The meteorology-chemistry interface processor (MCIP) for the CMAQ modeling system: updates through MCIPv3.4.1. Geosci. Model Dev., 3 (1), 243–256, doi:10.5194/gmd-3-243-2010. Pardo, L. H., 2010: Approaches for estimating critical loads of N and S deposition for forest ecosystems on U.S. federal lands. U.S. Department of Agriculture, Forest Service, https://doi.org/10.2737/NRS-GTR-71, 25 pp. Parra, R., 2018: Performance studies of planetary boundary layer schemes in WRFChem for the Andean region of southern Ecuador. Atmos. Pollut. Res., 9 (3), 411–428, doi:10.1016/j.apr.2017.11.011. 230 Peña-Arancibia, J. L., A. I. J. M. van Dijk, L. J. Renzullo, and M. Mulligan, 2013: Evaluation of precipitation estimation accuracy in reanalyses, satellite products, and an ensemble method for regions in Australia and South and East Asia. J. Hydrometeorol., 14 (4), 1323–1333, doi:10.1175/jhm-d-12-0132.1. Pérez, C., P. Jiménez, O. Jorba, M. Sicard, and J. M. Baldasano, 2006: Influence of the PBL scheme on high-resolution photochemical simulations in an urban coastal area over the western Mediterranean. Atmos. Environ., 40 (27), 5274–5297, doi:10.1016/ j.atmosenv.2006.04.039. Pernigotti, D., M. Gerboles, C. Belis, and P. Thunis, 2013: Model quality objectives based on measurement uncertainty. part II: NO2 and PM10. Atmos. Environ., 79, 869–878, doi:10.1016/j.atmosenv.2013.07.045. Pickard, D., P. de la Cueva Bueno, E. Olson, and C. Semmens, 2019: Evaluation of cumulative effects assessment methodologies for marine shipping. https://www.tc.gc.ca/en/services/marine/documents/EVALUATION_OF_ CUMULATIVE_EFFECTS_ASSESSMENT_METHODOLOGIES_FOR_MARINE_ SHIPPING_EN-2019-08-26.pdf. Pleim, J. E., 2007: A combined local and nonlocal closure model for the atmospheric boundary-layer. part i: model description and testing. J. Appl. Meteorol. Climatol., 46, 1383–1395, doi:10.1175/JAM2539.1. Pojar, J., K. Klinka, and D. A. Demarchi, 1991: Ecosystems of British Columbia. B.C. 231 Ministry of Forests, Victoria, 95–111 pp, https://www.for.gov.bc.ca/hfd/pubs/ docs/Srs/Srs06/chap6.pdf. Pontoppidan, M., J. Reuder, S. Mayer, and E. W. Kolstad, 2017: Downscaling an intense precipitation event in complex terrain: the importance of high grid resolution. Tellus A, 69 (1), 1271 561, doi:10.1080/16000870.2016.1271561. Porter, P. S., S. T. Rao, C. Hogrefe, E. Gego, and R. Mathur, 2015: Methods for reducing biases and errors in regional photochemical model outputs for use in emission reduction and exposure assessments. Atmos. Environ., 112, 178–188, doi: 10.1016/j.atmosenv.2015.04.039. Qiao, X., and Coauthors, 2015: Modeling dry and wet deposition of sulfate, nitrate, and ammonium ions in Jiuzhaigou National Nature Reserve, China using a sourceoriented CMAQ model: Part I. base case model results. Sci. Total Environ., 532, 831– 839, doi:10.1016/j.scitotenv.2015.05.108. Queen, A., and Y. Zhang, 2008: Examining the sensitivity of MM5–CMAQ predictions to explicit microphysics schemes and horizontal grid resolutions, part II—PM concentrations and wet deposition predictions. Atmos. Environ., 42 (16), 3856–3868, doi:10.1016/j.atmosenv.2007.12.066. Rakesh, P. T., R. Venkatesan, and C. V. Srinivas, 2013: Formulation of TKE based empirical diffusivity relations from turbulence measurements and incorporation in a lagrangian particle dispersion model. Environ. Fluid Mech., 13 (4), 353–369, doi: 10.1007/s10652-013-9273-8. 232 Reboredo, B., R. Arasa, and B. Codina, 2015: Evaluating sensitivity to different options and parameterizations of a coupled air quality modelling system over Bogotá, Colombia. Part I: WRF model configuration. Open J. Air Pollut., 4, 47–64, doi: 10.4236/ojap.2015.42006. Remesan, R., T. Bellerby, I. Holman, and L. Frostick, 2014: WRF model sensitivity to choice of parameterization: a study of the ‘York Flood 1999’. Theor. Appl. Climatol., 122 (1-2), 229–247, doi:10.1007/s00704-014-1282-0. Rife, D. L., C. A. Davis, and Y. Liu, 2004: Predictability of low-level winds by mesoscale meteorological models. Mon. Weather Rev., 132, 2553–2569, doi:10.1175/ MWR2801.1. Rio Tinto, 2013: So2 Technical Assessment volume 2. 450 pp, https: //www.riotinto.com/can/search/documents#main-search_e=0&main-search_ sxatags =bcworks%2Cenglish&canada-search_e=20. Rio Tinto, 2016: Rio tinto’s modernised kitimat aluminium smelter reaches full production. https://www.riotinto.com/news/releases/Kitimat-smelter-reaches-fullproduction. Rio Tinto, 2018: Annual environmental report. 43–45 pp, https:// www.riotinto.com/search/documents#main-search_e =0&main-search_sxatags= bcworks. Sakiyama, S., 2015: Guidelines for air quality dispersion modeling in British Columbia. https://www.bcogc.ca/node/13339/download. 233 Sarmadi, F., Y. Huang, G. Thompson, S. T. Siems, and M. J. Manton, 2019: Simulations of orographic precipitation in the Snowy Mountains of southeastern Australia. Atmos. Res., 219, 183–199, doi:10.1016/j.atmosres.2019.01.002. Schirmer, M., and B. Jamieson, 2015: Verification of analysed and forecasted winter precipitation in complex terrain. Cryosphere, 9 (2), 587–601, doi:10.5194/tc-9-5872015. Seinfeld, J. H., and S. N. Pandis, 2016: Atmospheric chemistry and physics : from air pollution to climate change. 3rd ed., John Wiley & Sons, Hoboken, New Jersey. Shafran, P., J. Woollen, W. Ebisuzaki, W. Shi, Y. Fan, R. Grumbine, and M. Fennessy, 2004: Observational data used for assimilation in the NCEP North American Regional Reanalysis. https://ams.confex.com/ams/84Annual/techprogram/paper_ 71689.htm. Shin, H. H., and S.-Y. Hong, 2015: Representation of the subgrid-scale turbulent transport in convective boundary layers at gray-zone resolutions. Mon. Weather Rev., 143, 250–271, doi:10.1175/MWR-D-14-00116.1. Skamarock, W. C., and Coauthors, 2019: A description of the advanced research WRF version 4. NCAR Tech. Note NCAR/TN-556+STR. 145, doi:10.5065/1dfh6p97. Smajla, I., D. K. Sedlar, B. Drljača, and L. Jukić, 2019: Fuel switch to LNG in heavy truck traffic. Energies, 12 (3), 515, doi:10.3390/en12030515. 234 Soil Classification Working Group, 1998: The Canadian System of Soil Classification (third edition). Agriculture and Agri–Food Canada. Publication 1646, 187 pp, http: //sis.agr.gc.ca/cansis/publications/manuals/1998-cssc-ed3/cssc3_manual.pdf. Solazzo, E., and S. Galmarini, 2016: Error apportionment for atmospheric chemistry-transport models — a new approach to model evaluation. Atmos. Chem. Phys., 16 (10), 6263–6283, doi:10.5194/acp-16-6263-2016. Srinivas, C., V. Yesubabu, D. H. Prasad, K. H. Prasad, M. Greeshma, R. Baskaran, and B. Venkatraman, 2018: Simulation of an extreme heavy rainfall event over Chennai, India using WRF: Sensitivity to grid resolution and boundary layer physics. Atmos. Res., 210, 66–82, doi:10.1016/j.atmosres.2018.04.014. Srinivas, C. V., K. B. R. R. H. Prasad, C. V. Naidu, R. Baskaran, and B. Venkatraman, 2015: Sensitivity analysis of atmospheric dispersion simulations by FLEXPART to the WRF-simulated meteorological predictions in a coastal environment. Pure Appl. Geophys., 173 (2), 675–700, doi:10.1007/s00024-015-1104-z. Stantec, 2015: 2014 vegetation inspection, monitoring and assessment program. 17 pp. Statistics Canada, 2018: Census profile, 2016 census. http://www12.statcan.gc.ca/ census-recensement/index-eng.cfm?GEOCODE=59#keystats. Sullivan, T. J., 2000: Aquatic Effects of Acidic Deposition. CRC Press, doi:10.1201/ 9780429187186. 235 Sun, J., H. Hu, Y. Li, L. Wang, Q. Zhou, and X. Huang, 2016: Effects and mechanism of acid rain on plant chloroplast ATP synthase. Environ. Sci. Pollut. Res., 23 (18), 18 296–18 306, doi:10.1007/s11356-016-7016-3. Sun, Q., C. Miao, Q. Duan, H. Ashouri, S. Sorooshian, and K.-L. Hsu, 2018: A review of global precipitation data sets: Data sources, estimation, and intercomparisons. Rev. Geophys., 56 (1), 79–107, doi:10.1002/2017rg000574. Sverdrup, H., and P. Warfvinge, 1995: Critical loads of acidity for Swedish forest ecosystems. Ecol. Bull., 44, 75–89, https://www.jstor.org/stable/20113151?seq= 2#metadata_info_tab_contents. Swall, J. L., and K. M. Foley, 2009: The impact of spatial correlation and incommensurability on model evaluation. Atmos. Environ., 43 (6), 1204–1217, doi: 10.1016/j.atmosenv.2008.10.057. Syrakov, D., M. Prodanova, E. Georgieva, I. Etropolska, and K. Slavov, 2016: Simulation of European air quality by WRF–CMAQ models using AQMEII-2 infrastructure. J. Comput. Appl. Math., 293, 232–245, doi:10.1016/j.cam.2015.01.032. Talbot, C., E. Bou-Zeid, and J. Smith, 2012: Nested mesoscale large-eddy simulations with WRF: performance in real test cases. J. Hydrometeorol., 13, 1421–1441, doi:10.1175/JHM-D-11-048.1. Tessum, C. W., J. D. Hill, and J. D. Marshall, 2015: Twelve-month, 12 km resolution North American WRF-Chem v3.4 air quality simulation: performance evaluation. Geosci. Model Dev., 8 (4), 957–973, doi:10.5194/gmd-8-957-2015. 236 Tewari, M. F., and Coauthors, 2004: Implementation and verification of the unified NOAH land surface model in the WRF model. Am. Meteorol. Soc., 11–15, https: //ams.confex.com/ams/84Annual/techprogram/paper_69061.htm. The Interior News, 2020: UPDATE: 11 arrested in day 3 of injunction enforcement against Wet’suwet’enl. https://www.interior-news.com/news/rcmp-continue- to-enforce-injunction-against-wetsuweten-opposing-coastal-gaslink/. Thompson, G., P. R. Field, R. M. Rasmussen, and W. D. Hall, 2008: Explicit forecasts of winter precipitation using an improved bulk microphysics scheme. part ii: implementation of a new snow parameterization. Mon. Weather Rev., 136, 5095–5115, doi:10.1175/2008MWR2387.1. Thunis, P., A. Pederzoli, and D. Pernigotti, 2012: Performance criteria to evaluate air quality modeling applications. Atmos. Environ., 59, 476–482, doi:10.1016/ j.atmosenv.2012.05.043. Tian, J., J. Liu, D. Yan, C. Li, and F. Yu, 2017: Numerical rainfall simulation with different spatial and temporal evenness by using a WRF multiphysics ensemble. Nat. Hazards Earth Syst. Sci., 17 (4), 563–579, doi:10.5194/nhess-17-563-2017. United Nations Economic Commission for Europe, 2017: Manual on methodologies and criteria for modelling and mapping critical loads and levels and air pollution effects, risks and trends. Chapter 3, https://www.umweltbundesamt.de/ sites/default/files/medien/4292/dokumente/ch3-mapman-2017-10.pdf. 237 United States Department of Agriculture, 2013: Conservation assessments for 5 species of lichens. 54–64, https://www.fs.fed.us/r6/sfpnw/issssp/documents2/ ca-li-5-lichens-2013-04.pdf. USEPA, 1996: Air quality criteria for particulate matter. https://nepis.epa.gov/ Exe/ZyPDF.cgi/20008MAJ.PDF?Dockey=20008MAJ.PDF, ePA/600/P-95/001aFcF. USEPA, 2009: Current methodologies in preparing mobile source port-related emission inventories. https://www.epa.gov/sites/production/files/2016-06/ documents/2009-port-inventory-guidance.pdf. USEPA, 2015: Interagency workgroup on air quality modeling phase 3 summary report: Long range transport and air quality related values. https://www3.epa.gov/ ttn/scram/11thmodconf/IWAQM3_LRT_Report-07152015.pdf. USEPA, 2016: List of designated reference and equivalent methods. https:// www3.epa.gov/ttnamti1/files/ambient/criteria/AMTICListDec2016-2.pdf. USEPA, 2017: Revisions to the guideline on air quality models: Enhancements to the AERMOD dispersion modeling system and incorporation of approaches to address ozone and fine particulate matter. https://www.govinfo.gov/content/pkg/ FR-2017-01-17/pdf/2016-31747.pdf. Vacek, S., and K. Matějka, 2010: Health status of forest stands on perma- nent research plots in the Krkonoše. Mts. J. Forest Sci., 56, 555–569, www.cabdirect.org/cabdirect/FullTextPDF/2011/20113038271.pdf. 238 https:// Veira, A., P. L. Jackson, B. Ainslie, and D. Fudge, 2013: Assessment of background particulate matter concentrations in small cities and rural locations—prince george, canada. J. Air Waste Manage. Assoc., 63 (7), 773–787, doi:10.1080/ 10962247.2013.789091. Vingarzan, R., 2007: Ambient particulate matter concentrations in Canada and background levels. Environment Canada, 20 pp, http://publications.gc.ca/ collections/collection_2007/ec/En56-215-2007E.pdf. Vivanco, M. G., and Coauthors, 2018: Modeled deposition of nitrogen and sulfur in Europe estimated by 14 air quality model systems: evaluation, effects of changes in emissions and implications for habitat protection. Atmos. Chem. Phys., 18 (14), 10 199–10 218, doi:10.5194/acp-18-10199-2018. Vorobeichik, E. L., M. R. Trubina, E. V. Khantemirova, and I. E. Bergman, 2014: Long-term dynamic of forest vegetation after reduction of copper smelter emissions. Russ. J. Ecol., 45, 498–162, doi:10.1134/S1067413614060150. Whitfield, C. J., J. N. Phelan, J. Buckley, C. M. Clark, S. Guthrie, and J. A. Lynch, 2018: Estimating base cation weathering rates in the USA: Challenges of uncertain soil mineralogy and specific surface area with applications of the PROFILE model. Water Air Soil Pollut., 229 (3), 15 pp, doi:10.1007/s11270-018-3691-7. WHO, 2000: Air Quality Guidelines for Europe (second edition). WHO Regional Publications, European Series, 91, 219–246 pp, http://www.euro.who.int/__data/ assets/pdf_file/0006/123099/AQG2ndEd_12ozonelevel.pdf?ua=1. 239 WHO, 2016: Ambient air pollution: a global assessment of exposure and burden of disease. ttps://www.who.int/phe/publications/air-pollution-global-assessment/en/. Williams, J. J., S. H. Chung, A. M. Johansen, B. K. Lamb, J. K. Vaughan, and M. Beutel, 2017: Evaluation of atmospheric nitrogen deposition model performance in the context of U.S. critical load assessments. Atmos. Environ., 150, 244–255, doi: 10.1016/j.atmosenv.2016.11.051. Williston, P. J., and Coauthors, 2016: Critical levels and loads and the regulation of industrial emissions in northwest British Columbia, Canada. Atmos. Environ., 146, 311–323, doi:10.1016/j.atmosenv.2016.08.058. Willmott, C., S. Robeson, and K. Matsuura, 2012: A refined index of model performance. Int. J. Climatol., 32, 2088–2094, doi:10.1002/joc.2419. Wong, J. S., S. Razavi, B. R. Bonsal, H. S. Wheater, and Z. E. Asong, 2017: Intercomparison of daily precipitation products for large-scale hydro-climatic applications over Canada. Hydrol. Earth Syst. Sci., 21 (4), 2163–2185, doi:10.5194/hess-212163-2017. Wyngaard, J., 2004: Toward numerical modeling in the “terra incognita. J. Atmos. Sci, 61, 1816–1826, doi:10.1175/1520-0469. Xie, B., J. Fung, A. Chan, and A. Lau, 2012: Evaluation of nonlocal and local planetary boundary layer schemes in the WRF model. J. Geophys. Res, 117, 12 103, doi: 10.1029/2011JD017080. 240 Yang, B., Y. Qian, G. Lin, R. Leung, and Y. Zhang, 2012: Some issues in uncertainty quantification and parameter tuning: a case study of convective parameterization scheme in the WRF regional climate model. Atmos. Chem. Phys., 12 (5), 2409–2427, doi:10.5194/acp-12-2409-2012. Yarwood, G., S. Rao, M. Yocke, and G. Whitten, 2005: Updates to the carbon bond chemical mechanism: CB05 final report to the USEPA. http://www.camx.com/ publ/pdfs/CB05_Final_Report_120805.pdf. Zhang, H., Z. Pu, and X. Zhang, 2013: Examination of errors in near-surface temperature and wind from WRF numerical simulations in regions of complex terrain. Wea. Forecasting, 28, 893–914, doi:10.1175/WAF-D-12-00109.1. Zhang, L., R. Vet, J. M. O'Brien, C. Mihele, Z. Liang, and A. Wiebe, 2009: Dry deposition of individual nitrogen species at eight Canadian rural sites. J. Geophys. Res., 114, D02 301, doi:10.1029/2008jd010640. 241 Appendices Appendix A Selection of meteorological year for model simulations To find a meteorological year that is representative of the study area, a ranking of differences between annual records for wind speed and precipitation in the TKV is presented. Wind speed is a major determinant of pollutant transport and dilution, while precipitation, which exceeds 2000 mm per annum in the Kitimat area, is a principal factor for the deposition of atmospheric contaminants. The ranking and selection of a meteorological year, thus, was based on these variables. Due to data availability for model evaluations, time span is years 2006–2018. Records are restricted to stations with longer and more complete archives. Mean wind speeds are hourly measurements while total precipitation are monthly sums of daily precipitation (rain + snow). See https://envistaweb.env.gov.bc.ca/DynamicTable2.aspx?G_ID= 327 and http://climate.weather.gc.ca/historical_data/search_historic_data_e.html for historical data. The computations are in two steps. First, pairwise absolute differences in monthly (Jan- 243 uary - December) values between all year pairs are calculated. Next, the mean of the differences for each year is calculated and ranked. Lower values are ranked higher (see Table A.1). Year 2017 is well ranked across sites for wind speed and is in top half of rank- Precipitation (mm) Wind speed (m s−1 ) Table A.1 The averages of differences between monthly means of wind speed and precipitation for each year (2006 to 2018) and their rankings. The less the difference, the better the ranking as a representative year. The empty values are instances of considerable missing data 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017 2018 3.90 4.39 4.01 3.89 4.47 3.92 4.81 4.35 4.37 4.56 5.34 3.58 3.78 4th 9th 6th 3rd 10th 5th 12th 7th 8th 11th 13th 1st 2nd 4.58 5.14 4.46 4.28 4.24 4.41 4.64 4.78 4.89 4.96 6.79 3.92 3.98 7th 12th 6th 4th 3rd 5th 8th 9th 10th 11th 13th 1st 2nd Terrace 6.79 7.54 6.79 8.43 7.34 8.06 7.04 8.45 8.43 7.96 8.61 6.98 7.48 Airport 1st 7th 1st 10th 5th 9th 4th 12th 10th 8th 13th 3rd 6th Kitimat 1343 1132 1035 1032 1019 1315 1281 - 1101 997 1213 1076 1079 (Town site) 12th 8th 4th 3rd 2nd 11th 10th - 7th 1st 9th 5th 6th - 568 559 499 518 625 549 602 580 548 700 524 557 - 8th 7th 1st 2nd 11th 5th 10th 9th 4th 12th 3rd 6th Haul road Whitesail Terrace ings for precipitation. Consequently, this year is selected as the meteorological year. Figure A.1 shows the profiles of this year’s values, and those of the averages for the 2006– 2018 period. Overall, the Kitimat 2017 data has -0.05 m s−1 , + 25.4 mm and - 0.8 ◦ C for mean wind speed, total precipitation and monthly air temperature, respectively, relative to the 2006–2018 values. 244 Figure A.1 Observed monthly average wind speed and precipitation in 2017 versus the averages for the 2006–2018 period. Terrace wind data is from the Airport that is 8 km south of the city 245 Appendix B WRF physics settings and model heights Table B.1 The under-listed physics and dynamics settings with option numbers (in WRF version 4.0) were used for generating meteorological fields. Physics/dynamics Option name Option number Land surface scheme Noah 2 Microphysics Thompson 8 Long- and short-wave radiation RRTMG 4 Cumulus Grell-Freitas 3 Diffusion option Full 2 Eddy coefficient option Horizontal Smagorinsky 4 246 Table B.2 Layer top heights (H1, H2, . . . , H39) for WRF simulations H1 24.9 m H11 1799.3 m H21 7699.1 m H31 14,628.7 m H2 81.7 m H12 2197.8 m H22 8414.1 m H32 15,320.0 m H3 153.8 m H13 2641.7 m H23 9117.5 m H33 16,010.7 m H4 245.0 m H14 3126.9 m H24 9810.3 m H34 16,700.8 m H5 359.6 m H15 3653.3 m H25 10,493.9 m H35 17,390.0 m H6 502.2 m H16 4224.1 m H26 11,174.4 m H36 18,079.0 m H7 677.7 m H17 4842.6 m H27 11,859.9 m H37 18,766.8 m H8 891.1 m H18 5512.2 m H28 12,551.5 m H38 19,455.4 m H9 1146.9 m H19 6233.1 m H29 13,244.6 m H39 20,145.0 m H10 1449.0 m H20 6233.1 m H30 8414.1 m 247 Appendix C Formulae for statistical measures Below are the formulas for the various statistical measures that are used in this dissertation: Mean bias (MB) = 1 n ( Mi − Oi ); n i∑ =1 Mean fractional bias (MFB) = Normalized mean bias (NMB) = MB O M−O 0.5( M + O) Normalized mean square error (NMSE)= Root mean square error (RMSE)= ( Mi − Oi )2 MO ( Mi − O i ) 2 Pearson’s correlation coefficient (r) = Modified index of agreement (IOA) = ∑( Mi − M)(Oi − O) ∑( Mi − M)2 ∑(Oi − O)2  n    ∑ | Mi − Oi |    i  1 − =1n ,      2 ∑ |Oi − O|  if n n i =1 i =1 n n i =1 i =1 ∑ | Mi − Oi | ≤ ∑ |Oi − O| i =1 n    2 ∑ |Oi − O|    i =1   − 1,  n     ∑ | Mi − Oi | if ∑ | Mi − Oi | > ∑ |Oi − O| i =1 Normalised standard deviation (NSD) = Standard deviation of modeled values σm = Standard deviation of observation data σo Where Mi and Oi are modeled and observed values (paired) respectively, n is the number of pairs and terms with overbars are mean values. 248 Appendix D Verification of atmospheric sounding Figure D.1 SkewT plot of simulations and observation for air temperature (solid lines), dew point (dashed lines), and wind for times in winter (top) and summer (bottom). Observation data are from the Annette Island radiosonde station, Alaska (latitude 55.03 ◦ N, longitude 131.57 ◦ W) 249 Appendix E Summer and winter days comparisons of air temperature and wind speed to seasonal averages Figure E.1a shows average hourly temperatures for July 22, 2017 and January 21, 2017, minus average hourly temperatures in summmer (left) and winter (right), respectively. The Light-gray portions between dark-gray backgrounds are daytime periods. The average daily temperature at Whitesail (Terrace) in summer is 15.4 ◦ C (16.1 ◦ C ). The average daily temperature at Whitesail (Terrace) in winter is -1.8◦ C (-2.0 ◦ C ). Figure E.1b shows seasonal windroses (S) versus those of the specified days (D) for daytime and nighttime periods. Marked on each plot is wind speed in m s−1 . Notice their overall correspondence for wind direction. 250 Figure E.1 Winterday (January 21, 2017) and summerday (July 22, 2017 ) comparisons to observed seasonal averages of air temperature and wind. 251 Appendix F Performance statistics for simulation of surface meteorological variables 252 Table F.1 Statistical performance of simulations of air temperature at a diurnal time scale for each season. Values in bold indicate that criteria (refer to Table 2.3) are met, and italicized means a value is top-ranked for a station. An instance where more than two schemes are joint-best is not ranked. Fitness of individual schemes are assessed from counts of top-ranked and satisfactory evaluations (bolded and italicized). Nanakwa, Whitesail, Onion Lake and Terrace stations are N, W, O and T, respectively. spring Measure summer autumn winter PBL scheme N W O T N W O T N W O T N W O T MYJ -0.47 -1.89 -1.54 -2.37 -0.96 -0.35 -0.32 -0.41 1.14 0.64 3.80 0.93 1.99 0.18 0.33 0.55 MYNN3 0.06 -1.65 -1.28 -2.19 -0.70 -0.39 -0.42 -0.53 1.16 0.99 4.13 1.18 2.17 0.50 0.75 0.89 UW 0.04 -1.71 -1.36 -2.23 -0.66 -0.18 -0.18 -0.27 1.18 0.93 4.10 1.17 2.12 0.32 0.48 0.69 ACM2 -0.03 -1.92 -1.72 -2.38 -0.44 -0.40 -0.46 -0.47 1.15 0.65 3.65 0.93 1.63 -0.05 0.08 0.24 YSU -0.13 -2.06 -1.86 -2.49 -0.50 -0.51 -0.56 -0.58 1.06 0.45 3.52 0.80 1.47 -0.28 -0.07 0.09 SH -0.12 -2.05 -1.88 -2.49 -0.50 -0.51 -0.57 -0.58 1.07 0.45 3.49 0.81 1.48 -0.27 -0.08 0.09 MYJ 1.35 2.25 1.92 2.63 1.72 0.94 1.22 1.19 1.44 1.09 3.82 1.20 2.08 0.77 0.58 0.77 MYNN3 1.25 2.17 1.84 2.53 1.54 1.09 1.40 1.32 1.44 1.33 4.15 1.42 2.24 0.94 0.91 1.06 UW 1.27 2.12 1.81 2.53 1.57 0.94 1.24 1.16 1.48 1.27 4.13 1.39 2.21 0.81 0.68 0.88 ACM2 1.25 2.31 2.13 2.66 1.52 1.05 1.38 1.27 1.48 1.22 3.76 1.28 1.74 0.96 0.55 0.81 YSU 1.27 2.41 2.26 2.77 1.52 1.13 1.44 1.33 1.40 1.16 3.64 1.20 1.58 1.03 0.61 0.85 SH 1.28 2.42 2.27 2.76 1.54 1.12 1.44 1.34 1.41 1.16 3.62 1.20 1.59 1.05 0.65 0.84 MYJ 0.06 0.96 0.94 0.92 -0.10 0.97 0.94 0.93 -0.22 0.90 0.94 0.90 -0.08 0.89 0.96 0.87 MYNN3 0.06 0.97 0.95 0.93 -0.07 0.98 0.95 0.94 -0.19 0.91 0.94 0.90 -0.10 0.89 0.96 0.87 UW 0.04 0.96 0.94 0.92 -0.12 0.97 0.94 0.93 -0.22 0.90 0.93 0.89 -0.13 0.88 0.95 0.86 ACM2 0.12 0.92 0.88 0.89 -0.11 0.95 0.91 0.91 -0.24 0.84 0.87 0.85 -0.05 0.71 0.88 0.71 YSU 0.10 0.91 0.86 0.88 -0.09 0.94 0.90 0.91 -0.21 0.83 0.86 0.84 0.01 0.70 0.86 0.66 SH 0.09 0.91 0.86 0.88 -0.11 0.94 0.90 0.91 -0.22 0.83 0.86 0.85 0.00 0.69 0.85 0.67 MYJ 0.43 0.61 0.65 0.45 0.25 0.84 0.80 0.81 -0.05 0.70 -0.42 0.63 -0.56 0.70 0.75 0.62 MYNN3 0.47 0.65 0.67 0.49 0.32 0.81 0.77 0.78 -0.06 0.63 -0.47 0.56 -0.59 0.62 0.61 0.46 UW 0.47 0.65 0.67 0.48 0.30 0.83 0.80 0.80 -0.08 0.65 -0.47 0.57 -0.58 0.68 0.72 0.56 ACM2 0.47 0.59 0.59 0.44 0.30 0.82 0.78 0.78 -0.05 0.70 -0.40 0.64 -0.46 0.67 0.77 0.65 YSU 0.46 0.57 0.56 0.42 0.31 0.80 0.77 0.77 0.02 0.73 -0.38 0.67 -0.40 0.64 0.75 0.62 SH 0.45 0.57 0.56 0.42 0.30 0.80 0.76 0.77 0.01 0.73 -0.37 0.67 -0.40 0.64 0.74 0.63 MB (K) RMSE (K) r IOA 253 Table F.2 Performance statistics for specific humidity at diurnal time scale for each season are given below. Values in bold indicate that criteria (refer to Table 2.3) are met and italicized means a value is top-ranked for each station. An instance where more than two schemes are joint-best is not ranked. Fitness of individual schemes are assessed from counts of top-ranked and satisfactory evaluations (bold and italicized). Onion Lake and Terrace stations are O and T, respectively. spring summer autumn winter O O O spring summer autumn winter O O O T O T PBL scheme O T T T T T T MB (g kg−1 ) r MYJ -0.47 -0.19 -0.27 -0.05 -0.87 -0.61 -0.37 -0.05 0.79 0.82 0.71 0.76 0.81 0.90 0.84 0.85 MYNN3 -0.43 -0.16 -0.19 0.03 -0.86 -0.6 -0.35 -0.03 0.78 0.81 0.67 0.70 0.81 0.90 0.84 0.85 UW -0.54 -0.25 -0.43 -0.18 -0.94 -0.67 -0.37 -0.05 0.78 0.81 0.71 0.76 0.81 0.90 0.84 0.85 ACM2 -0.71 -0.39 -0.66 -0.41 -1.09 -0.82 -0.48 -0.14 0.77 0.81 0.75 0.79 0.77 0.89 0.82 0.84 YSU -0.69 -0.41 -0.68 -0.48 -1.01 -0.82 -0.48 -0.14 0.79 0.82 0.74 0.78 0.78 0.89 0.84 0.84 SH -0.65 -0.33 -0.53 -0.27 -1.06 -0.78 -0.5 -0.14 0.76 0.81 0.72 0.77 0.78 0.90 0.83 0.84 RMSE (g kg−1 ) IOA MYJ 1.06 0.91 1.25 1.07 1.26 1.17 0.89 0.79 0.56 0.66 0.60 0.65 0.54 0.73 0.72 0.76 MYNN3 1.06 0.93 1.34 1.20 1.26 1.16 0.88 0.79 0.56 0.66 0.57 0.62 0.55 0.74 0.72 0.76 UW 1.08 0.91 1.27 1.06 1.32 1.19 0.90 0.79 0.54 0.66 0.58 0.65 0.52 0.73 0.71 0.76 ACM2 1.02 0.96 1.27 1.03 1.48 1.31 1.01 0.85 0.48 0.63 0.56 0.64 0.46 0.70 0.67 0.74 YSU 1.13 0.94 1.28 1.07 1.40 1.31 0.95 0.83 0.50 0.64 0.55 0.63 0.49 0.69 0.69 0.75 SH 1.19 0.96 1.29 1.05 1.46 1.27 1.01 0.85 0.50 0.64 0.57 0.65 0.47 0.71 0.68 0.74 254 Table F.3 Wind direction bias per season (in ◦ ) using all hourly values. Values >0 ◦ signify clockwise bias. Values in bold indicate that being within ± 10 ◦ is achieve and italicized means a value is top-ranked for each station. An instance where more than two schemes are joint-best is not ranked. Fitness of individual schemes are assessed from counts of top-ranked and satisfactory evaluations (bold and italicized). Nanakwa, Whitesail, Onion Lake and Terrace stations are N, W, O and T, respectively. spring summer autumn winter PBL scheme N W O T N W O T N W O T N W O T MYJ 10.5 16.9 -31.8 10.9 3.7 20.5 -23.7 0.4 2.6 20.4 -40.1 22.9 1.9 17.0 -43.8 26.8 MYNN3 5.7 18.3 -38.6 9.2 4.8 18.8 -29.8 -4.9 3.4 20.7 -43.2 17.8 0.1 14.6 -48.8 25.3 UW 5.2 16.2 -33.2 11.2 3.5 19.1 -25.3 -2.3 0.6 21.8 -37.8 20.5 -1.6 17.2 -45.6 22.4 ACM2 3.2 21.7 -26.5 15.2 3.9 21.8 -19.4 0.9 1.4 24.5 -34.9 21.1 -10.0 20.9 -43.8 19.1 YSU 5.2 20.6 -33.9 11.6 6.3 21.6 -25.0 -0.8 1.2 21.5 -40.7 19.1 -9.1 17.0 -52.3 17.6 SH 5.9 19.6 -32.8 11.2 7.0 21.6 -24.9 -1.6 1.7 21.0 -40.4 19.2 -9.4 17 -53.5 18.4 255 Table F.4 Performance statistics for wind speed at diurnal time scale for each season is presented in the table below. Values in bold indicate that criteria (refer to Table 2.3) are met and italicized means a value is top-ranked for a station. An instance where more than two schemes are joint-best is not ranked. Fitness of individual schemes are assessed from counts of top-ranked and satisfactory evaluations (bold and italicized). Nanakwa, Whitesail, Onion Lake and Terrace stations are N, W, O and T, respectively. MB (m s −1 ) RMSE (m s−1 ) PBL spring summer scheme N W O T N MYJ 1.39 0.98 0.71 2.64 0.46 MYNN3 1.53 0.87 0.50 2.37 UW 1.64 0.99 0.68 ACM2 1.98 0.99 YSU 1.91 SH autumn W winter O T N W O T N W O T 1.4 0.26 2.43 1.21 1.63 1.85 3.02 0.81 1.32 1.84 3.67 0.56 1.14 -0.14 1.88 1.21 1.57 1.82 2.80 0.83 1.29 1.82 3.50 2.60 0.71 1.50 0.34 2.43 1.24 1.67 1.92 3.01 0.79 1.25 1.77 3.55 0.50 2.45 0.83 0.88 -0.13 1.91 1.77 1.47 1.67 2.78 1.61 1.75 1.97 3.73 0.88 0.41 2.31 0.77 0.76 -0.21 1.79 1.72 1.31 1.58 2.63 1.65 1.63 1.98 3.53 1.95 0.89 0.45 2.31 0.88 0.78 -0.18 1.85 1.77 1.33 1.59 2.69 1.66 1.67 2.02 3.52 MYJ 1.48 1.00 0.94 2.65 1.06 1.43 0.72 2.44 1.27 1.65 1.86 3.03 1.01 1.35 1.90 3.68 MYNN3 1.58 0.91 0.97 2.41 0.91 1.17 0.97 1.91 1.28 1.59 1.82 2.81 1.01 1.31 1.87 3.51 UW 1.74 1.01 0.92 2.62 1.18 1.53 0.79 2.43 1.29 1.71 1.94 3.01 0.98 1.28 1.85 3.55 ACM2 2.02 1.06 1.01 2.50 1.04 0.97 1.16 1.94 1.85 1.48 1.68 2.79 1.73 1.78 2.05 3.74 YSU 1.95 0.95 0.94 2.36 0.99 0.84 1.10 1.81 1.80 1.32 1.59 2.64 1.76 1.66 2.06 3.54 SH 2.00 0.97 0.91 2.36 1.11 0.84 1.08 1.87 1.84 1.34 1.60 2.69 1.77 1.70 2.09 3.53 MYJ 0.35 0.95 0.95 0.89 -0.01 0.96 0.98 0.98 0.24 0.94 0.91 0.74 -0.44 0.51 0.56 -0.09 MYNN3 0.56 0.95 0.94 0.53 0.19 0.93 0.97 0.94 0.04 0.94 0.93 0.54 -0.48 0.44 0.59 0.13 UW 0.33 0.95 0.97 0.87 0.01 0.95 0.98 0.99 0.37 0.94 0.91 0.75 -0.55 0.46 0.34 0.22 ACM2 0.54 0.90 0.94 0.13 0.44 0.94 0.95 0.92 -0.53 0.90 0.88 0.16 -0.51 -0.14 0.02 0.13 YSU 0.60 0.92 0.95 0.42 0.43 0.92 0.98 0.97 -0.63 0.88 0.87 0.33 -0.59 -0.03 0.24 0.06 SH 0.50 0.90 0.96 0.56 0.3 0.94 0.98 0.98 -0.50 0.89 0.88 0.54 -0.55 -0.10 0.33 0.14 MYJ -0.38 0.16 0.61 -0.64 0.30 -0.04 0.78 -0.51 -0.45 -0.68 -0.68 -0.88 -0.46 -0.77 -0.51 -0.96 MYNN3 -0.44 0.25 0.58 -0.6 0.35 0.15 0.69 -0.37 -0.45 -0.67 -0.68 -0.87 -0.46 -0.76 -0.50 -0.96 UW -0.47 0.15 0.61 -0.64 0.22 -0.10 0.75 -0.51 -0.46 -0.69 -0.69 -0.88 -0.44 -0.76 -0.49 -0.96 ACM2 -0.56 0.16 0.56 -0.61 0.28 0.34 0.62 -0.38 -0.62 -0.65 -0.65 -0.87 -0.69 -0.83 -0.54 -0.96 YSU -0.55 0.25 0.59 -0.59 0.31 0.43 0.65 -0.34 -0.61 -0.60 -0.63 -0.86 -0.70 -0.81 -0.54 -0.96 SH -0.56 0.24 0.61 -0.59 0.23 0.42 0.65 -0.36 -0.62 -0.61 -0.63 -0.86 -0.70 -0.82 -0.55 -0.96 r IOA 256 Appendix G Centreline profile profiles of key WRF output variables 257 Figure G.1 Below are are profiles of WRF diagnosed variables along-valley and through the coastal inlet (see Fig. 2.1). PBLH is planetary boundary-layer height, CLDB is cloud base height, both in meters above terrain height. CFRAC is cloud fraction. 258 Appendix H Formulae for indices of precipitation prediction Precipitation forecast scores are from counts of true and false responses for predictions and observation in the form of a contingency table: Such that: Table H.1 Contingency table for precipitation events forecast Modeled True (T) False (F) True (T) TT TF False (F) FT FF Observation Probability of detection (POD) = True positives TT = True positives + False negatives TT + TF Frequency alarm ratio (FAR) = False positives FT = True positives + False positives TT + FT Equitable threats score (ETS) = True positives -C TT - C = True positives + False positives + False negatives TT + TF + FT where C = (TT + TF)(TT + FT ) (TT + TF)(FT + FF ) Frequency bias index (FBI) = True positives + False positives TT + FT = True positives + False negatives TT + TF 259 Appendix I The implication of sulfur exceedance on vegetated land From Table 7.3, average S deposition exceedance is ∼ 35 kg ha yr −1 1 km2 = 100 ha A 10 km2 area (the lower estimate) would recieve 10 x 100 ha x 35 kg kg = 35000 = 35 ha yr yr tonnes yr−1 Volume 2, page 14 of the technical assessment of Rio Tinto’s application for ammendment of its SO2 emission permit (Rio Tinto 2013), notes 35 years of future aluminum production, hence total excess surfur deposition of: 35 yr x 35 tonnes yr−1 = 1225 tonnes. For a 6 m distance between trees, and 5-6m between tree rows (Food and Agricultural Organization 1989), I ha can contain 333–400 trees. Therefore, a 10 km2 area (1000 ha) can contain 1000x the above or 333,000–400,000 trees. 260