theory and equations: supplementary material table 1...

13
Oceanic macromolecular surfactant estimates 1 Theory and equations: supplementary material 1 2 Our demonstration of marine surfactant/organic variability requires the application of 3 macromolecular physical chemistry concepts at divergent levels. Some readers will 4 already be comfortable with the full range of logic, but others will be more familiar with 5 either its marine or atmospheric aspects. In order to unify the concepts, we present an 6 equation list here in appendix form. The exposition is structured as an ocean-up narrative, 7 describing organic behaviors for major water-air interface types and the nascent aerosol. 8 Chemical reference points and indicators cited in the text tables are completely defined. 9 In order of appearance the quantities under consideration are: 10 11 Table 1 (general surface activity): Half saturation concentration C 1/2 (moles carbon per 12 liter) and the maximum excess or surface concentration Γ max (carbon atoms per square 13 angstrom with conversion to moles carbon per square meter as needed) 14 15 Table 2 (ocean surfactants): Normalized bulk concentration C/C 1/2 (dimensionless) and 16 the closely related relative excess measure Γ max C/C 1/2 (moles carbon per square meter) 17 18 Table 3 (ocean to atmosphere transition): two dimensional phase state analogs (solid 19 versus expanded films etc.) 20 21 Table 3 (atmospheric surfactants): Driving force for spray coverage 1 Molar/C 1/2 22 (dimensionless), surface to aqueous carbon ratios (Γ max /1 Molar)(A/V)( also 23 dimensionless), the surface pressure Π which is also a tension reduction (mJ/m 2 ) 24 25 Table 3 (atmospheric macromolecular chemistry): Parameterized hygroscopicity κ 26 (dimensionless), organic to carbon mass ratio (dimensionless) and individual compound 27 density (g/cc). 28 29 For practical purposes, we restrict our thinking to a thin but global band embracing the 30 oceanic upper mixed layer, the sea surface and the immediate vicinity of sea spray 31 droplets. Conservation of free energy is a useful starting point in all situations. But it 32 must be conceived in the multicomponent multiphase sense -between solvent water, high 33 molecular weight organics and air all taken in the context of bubbles, the global ocean 34 boundary and finally the nascent aerosol. We adapt derivations primarily from two 35 prominent text book series –Adamson (1960) to Adamson and Gast (1997) and Seinfeld 36 (1986) to Seinfeld and Pandis (2006) for surfactant and aerosol chemical issues 37 respectively. An historical view is attempted, with reference to seminal research in 38 several relevant fields. For example the classic physical chemical contributions of Gibbs 39 and Langmuir are acknowledged, but simultaneously we draw on early and rapidly 40 evolving environmental studies of the marine aerosol organics. 41 42 Pointers have been inserted at several appropriate locations in the main text, directing 43 readers to the equations and theory. A more complete development is available from the 44 COSIM group at Los Alamos (Climate Ocean Sea Ice Modeling) as a technical report in 45

Upload: ngophuc

Post on 10-Mar-2018

217 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  1  

Theory and equations: supplementary material 1   2  Our demonstration of marine surfactant/organic variability requires the application of 3  macromolecular physical chemistry concepts at divergent levels. Some readers will 4  already be comfortable with the full range of logic, but others will be more familiar with 5  either its marine or atmospheric aspects. In order to unify the concepts, we present an 6  equation list here in appendix form. The exposition is structured as an ocean-up narrative, 7  describing organic behaviors for major water-air interface types and the nascent aerosol. 8  Chemical reference points and indicators cited in the text tables are completely defined. 9  In order of appearance the quantities under consideration are: 10   11  Table 1 (general surface activity): Half saturation concentration C1/2 (moles carbon per 12  liter) and the maximum excess or surface concentration Γmax (carbon atoms per square 13  angstrom with conversion to moles carbon per square meter as needed) 14   15  Table 2 (ocean surfactants): Normalized bulk concentration C/C1/2 (dimensionless) and 16  the closely related relative excess measure Γmax C/C1/2 (moles carbon per square meter) 17   18  Table 3 (ocean to atmosphere transition): two dimensional phase state analogs (solid 19  versus expanded films etc.) 20   21  Table 3 (atmospheric surfactants): Driving force for spray coverage 1 Molar/C1/2 22  (dimensionless), surface to aqueous carbon ratios (Γmax/1 Molar)(A/V)( also 23  dimensionless), the surface pressure Π which is also a tension reduction (mJ/m2) 24   25  Table 3 (atmospheric macromolecular chemistry): Parameterized hygroscopicity κ 26  (dimensionless), organic to carbon mass ratio (dimensionless) and individual compound 27  density (g/cc). 28   29  For practical purposes, we restrict our thinking to a thin but global band embracing the 30  oceanic upper mixed layer, the sea surface and the immediate vicinity of sea spray 31  droplets. Conservation of free energy is a useful starting point in all situations. But it 32  must be conceived in the multicomponent multiphase sense -between solvent water, high 33  molecular weight organics and air all taken in the context of bubbles, the global ocean 34  boundary and finally the nascent aerosol. We adapt derivations primarily from two 35  prominent text book series –Adamson (1960) to Adamson and Gast (1997) and Seinfeld 36  (1986) to Seinfeld and Pandis (2006) for surfactant and aerosol chemical issues 37  respectively. An historical view is attempted, with reference to seminal research in 38  several relevant fields. For example the classic physical chemical contributions of Gibbs 39  and Langmuir are acknowledged, but simultaneously we draw on early and rapidly 40  evolving environmental studies of the marine aerosol organics. 41   42  Pointers have been inserted at several appropriate locations in the main text, directing 43  readers to the equations and theory. A more complete development is available from the 44  COSIM group at Los Alamos (Climate Ocean Sea Ice Modeling) as a technical report in 45  

Page 2: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  2  

the LAUR-LAMS series. Global mapping exercises and initial extensions to full, 46  competitive Langmuir equilibria are currently under review in Burrows et al (2014). 47   48  Table 1 (surface activity): C1/2 and Γmax 49   50  To establish our first indicator quantities, we take the point of view of a marine 51  macromolecule or polymer, possibly surface active, generated by phytoplanktonic cell 52  disruption in upper layers of the global ocean then transported toward or into the 53  atmosphere (schematics in Blanchard 1963 and 1989, Russell et al 2010). Within the 54  water column, organic structures are of course subject to an intense reactive chemistry -55  enzymatic and photolytic degradation plus recombination (Benner, 2002; Hansell et al 56  2012). But in the upper few meters they may also be confronted with aerobic interfaces, 57  associated with bubble plumes injected from above (Hoffman and Duce, 1976; Cunliffe 58  et al 2011). The wave-generated bubble distribution peaks between 30 and 300 microns 59  (Woolf 1997). At such sizes curvature may be neglected (Clift et al 1978), which 60  significantly simplifies our task at the outset. 61   62  To construct a basic thermochemical model, imagine a generic system consisting of 63  biomacromolecules and sea salt as solutes, solvent water surrounding them, a subset of 64  organics residing at the interface and then an overlying gas phase. Given bulk liquid and 65  gaseous mixtures of i components, total moles and Gibbs free energy may be expressed 66  as simple sums. But the surfactant reservoir is also accounted explicitly, and 67  rearrangement emphasizes its centrality. 68   69  𝑛!! + 𝑛!

! +  𝑛!!"#$%&' = 𝑛!;  𝐺! + 𝐺! + 𝐺!"#$%&' = 𝐺 (1) 70  

 𝑛!!"#$%&' = 𝑛! −  𝑛!! + 𝑛!

! ;  𝐺!"#$%&' = 𝐺 − 𝐺! + 𝐺! (2) 71   72  A general incremental change in free energy is just 73   74  𝑑𝐺 = −𝑆𝑑𝑇 + 𝑉𝑑𝑝 + 𝜎𝑑𝐴 + 𝜇!! 𝑑𝑛! (3) 75   76  where σ is the surface tension in units, for example, of N/m or J/m2 and other symbols 77  have their usual meanings. This is simply a textbook dG given addition of the surface 78  energy contribution. Under local conditions in the marine aqueous system, temperature 79  may be considered constant. Furthermore the volume of the dividing plane is zero by 80  definition (Adamson and Gast 1997, Tuckermann 2007). Since all chemical potentials are 81  in balance at equilibrium, the expression for surface free energy change simplifies and is 82  readily integrated. 83   84  𝜇!!"#$%&' = 𝜇!! = 𝜇!

! (4) 85  𝑑𝐺!"#$%&' = 𝜎𝑑𝐴 + 𝜇!! 𝑑𝑛!

!"#$%&' (5) 86  𝐺!"#$%&' = 𝜎𝐴 + 𝜇!! 𝑛!

!"#$%&' (6) 87   88  

Page 3: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  3  

We seek to understand tension as a function of composition. Application of the product 89  rule yields the expansion in 7. Gibbs Duhem-type arguments and matching with 5 then 90  provide the key relationship. 91   92  𝑑𝐺!"#$!"# = 𝜎𝑑𝐴 + 𝐴𝑑𝜎 + 𝑛!

!"#$%&'! 𝑑µμ! +   𝜇!! 𝑑𝑛!

!"#$%&' (7) 93  𝐴𝑑𝜎 = − 𝑛!

!"#$%&'! 𝑑µμ! (8) 94  

95  At this point we focus for purposes of clarity on a single surface active organic. The 96  summation is obviated and it is natural to convert moles into a two dimensional 97  concentration, Γ = n/A. Typically this quantity is referred to as an “excess” of the 98  compound, because it is not present in the aqueous phase. For any dilute solute, chemical 99  potential is related to the logarithm of concentration C. Various forms of the Gibbs 100  surface equation therefore obtain: 101   102  𝑑𝜎 = −Γ𝑑𝜇;𝑑𝜇 = 𝑅𝑇𝑑𝑙𝑛𝐶;  𝑑𝜎 𝑑𝐶 = − Γ𝑅𝑇 𝐶   (9) 103   104  Now suppose that in a weak solution linearity is in force for the above relationship. 105  Macromolecules necessarily reduce the interfacial tension. Surface pressure Π is 106  typically defined as the difference in σ manifested by pure and water-solute situations. 107   108  𝜎 = 𝜎∗ +−𝑐𝑜𝑛𝑠𝑡.𝐶;  Π = 𝑐𝑜𝑛𝑠𝑡.C;  𝑑𝜎 𝑑𝐶 = −𝑐𝑜𝑛𝑠𝑡. ;  Π = Γ𝑅𝑇 (10) 109   110  The final expression 10 is sometimes referred to as the “two dimensional ideal gas law” 111  since back-substitution of the excess yields ΠA = nRT. In principle, some combination of 112  9 with the simple “law” might be used to model the behavior of hypothetical surfactants 113  in ocean/aerosol systems. But in fact the idealized relationships often break down under 114  ambient conditions. They refer to point solutes/adsorbers lacking structure and molecular 115  interaction. Real systems exhibit repulsions whether in the bulk or a concentrating film. 116  Polymers experience denaturing and wrapping in solution (Lipitov and Sergeeva 1984, 117  Birdi, 1989). Hydrophobic groups may tilt and rotate on the surface. It has long been 118  recognized that protein and lipid series begin to deviate from ideality as early as 10-5 and 119  10-3 J/m2. (Schofield and Rideal 1925, Guastalla 1939, Ter Minassian-Saraga 1956). And 120  in the present work we do not go so far as to consider two dimensional mixtures. 121   122  The Gibbs equation couples neatly to results from Langmuir adsorption kinetics to yield a 123  fit to the tension curve in the region of mounting surfactant influence. Rates of site 124  occupation and subsequent desorption are given by ka C, where the proportionality 125  constant has units of per (molar time), and kd in the reverse direction which is first order. 126  Fractional coverage is then a function of the ratio between the two, which in turn may be 127  thought of as an adsorptive equilibrium constant. The interested reader can readily 128  demonstrate that its reciprocal is simply the concentration at half coverage C1/2. 129   130  𝐾 = 𝑘! 𝑘! ;  𝜃 = 𝐾𝐶 1+ 𝐾𝐶 ;  𝐶! ! = 𝐾!! (11) 131   132  

Page 4: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  4  

An assymptote for monolayer coverage Γmax may also be assumed, although even this is 133  not a given as polymers reconfigure, stack and react (Ter Minassian-Saraga 1956, 134  McGregor and Barnes 1978, Graham and Phillips 1979a through c). A particularly 135  compact mathematical form is then attained. Substitution and integration yield what is 136  often called the Langmuir-Szyszkowski equation (LS), after major contributors working 137  near the beginning of the last century. 138   139  Γ = Γ!"#𝜃 = Γ!"#𝐾𝐶 1+ 𝐾𝐶 (12) 140  𝑑𝜎 = − Γ𝑅𝑇𝑑𝐶 𝐶 ;  𝑑𝜎 = −Γ!"#𝐾𝑅𝑇 𝑑𝐶/ 1+ 𝐾𝐶 (13) 141  Π = Γ!"#𝑅𝑇𝑙𝑛 1+ 𝐾𝐶 (14) 142   143  Equilibrium constants C1/2 in text Table 1 were obtained by fitting laboratory surface 144  tension or pressure data to the Gibbs equation in forms 13 and 14, with one significant 145  exception. Stearic acid is sufficiently insoluble in seawater that a kinetic dissolution 146  approach had to be applied (Ter Minassian-Seraga 1956, Brzozowska et al 2012) . The 147  limiting 2D concentrations Γmax are measurable directly, by weighing out a particular 148  sample then generating the spread film on a flat water surface (Graham and Phillips 149  1979b). Values are often reported in mg/m2 but we convert to atomic or mole units for 150  convenience, first in visualizing the microscopic system then conducting comparisons 151  with the bulk. 152   153  Note that in a true Langmuir-kinetic situation, temporal dependencies are strongly 154  implied (Graham and Phillips 1979a, Lipitov and Sergeeva 1984, Adamson and Gast 155  1997). The rate constants ka will enfold differing polymeric diffusion rates, and 156  reciprocal kd are a set of desorption time scales depending on binding energies at diverse 157  sites. The potential for disequilibrium is real but is bypassed intentionally in our study. 158  We plan to address this point as it becomes possible and necessary to conduct simulations 159  at the molecular scale. 160   161  Table 2 (ocean surfactants): C/C1/2 and Γmax C/C1/2 162   163  If the distinctions “i” were maintained and propagated beyond 8 in order to represent 164  multiple adsorbing agents, our manipulations would reflect competition for the available 165  bubble surface area (Burrows et al 2014). Multicomponent coverage takes the chemically 166  resolved form 167   168  𝜃! = 𝐾!𝐶! 1+   𝐾!! 𝐶! ;  𝜃!"!#$ = 𝜃!!     (15) 169   170  The derivation once again involves the kinetics of approach to (and desorption from) a 171  water-air interface. It is given in many physical chemistry source books (e.g. Adamson, 172  1960, Laider 1965). In the context of the simultaneous equilibria of 15, an individual 173  ratio KjCj = C/C1/2 for the jth species provides a quick assessment of relative coverage. 174  This is our strategy as a preliminary analysis method, and the same arguments apply to 175  equation 12. Therefore the quantity ΓmaxC/C1/2 can be thought of as a mass weighted 176  coverage analog. The usual caveats with regard to nonlineary and nonideality must 177  naturally be kept in mind. The surfactant systems are in reality structurally complex and 178  

Page 5: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  5  

dynamic, entailing polymer conformation shifts, molecular interactions, chemical 179  reactions and more so that Gibbs-Langmuir is really only an approximation. 180   181  Table 3 (transition to atmosphere): The surfactant phase state 182   183  As the bubble field advances through the water column carrying long chain carbon 184  upward, a global ocean-atmosphere boundary is approached which is physically quite 185  heterogeneous (Liss 1975, Hardy 1982, Cunliffe et al 2011). Photochemistry, gels, foams, 186  organisms and even entire ecosystems are supported there. Organic pollutants accumulate 187  because they too tend to be surface active. Particles are continually deposited from the 188  atmosphere and contribute to the mix. But a true aqueous salt system is interspersed, and 189  this particular portion of the interface is entirely amenable to the current logic. We find 190  ourselves firmly in the realm of Σini

surface in equations 4 through 6. In the real ocean 191  vertical microchemical gradients must exist just below the atmosphere, and such factors 192  can ultimately be parameterized. For present purposes however, we make the zeroth order 193  assumption that two dimensional ocean-atmosphere chemistry resembles that of the 194  approaching bubbles. 195   196  As the excess Γ is increased during a given laboratory experiment, planar versions of 197  familiar “phase transitions” are documented. Order may accumulate as the adsorbing 198  entities align head to tail or effectively crystallize within a monolayer. Analogs of three 199  dimensional liquid and solid states appear as kinks in surface pressure versus area 200  experiments (Christodoulou and Rosano 1968, Liss 1975, Frew 1997, Donaldson and 201  Vaida 2006; Brozozowska et al 2012). Early work in the marine realm invoked the Van 202  der Waals equation of state as a framework for comprehending such phase phenomenon 203  in coastal waters (Barger and Means 1985). This is one of the ways in which oceanic 204  microlayer surfactants were first shown to be imperfect. In fact their behavior is often 205  nonsolid and expanded but not chaotic enough to be referred to as “gaseous” –the level of 206  disorder is intermediate on the Gibbs plane. Hence the classifications adopted in Table 3. 207   208  The film state is partially responsible for uncertainties in species-dependent sea-air gas 209  transfer. Multiple mechanisms are at play. Macromolecules introduce purely mechanical 210  barriers to the mobility of small molecules (Clift et al 1976, Leifer and Patro 2002). They 211  also alter viscoelastic properties of the laminar layer across which molecular diffusion 212  must take place (Frew et al 1990). In overlap with (and/or feedback onto) the 213  fundamental arguments here, surfactants distributed along the integrated bubble field 214  modulate dissolution and therefore the Henry’s Law augmentation to gas transport (Frew 215  1997). At high wind speed, this is the strongest atmospheric reintroduction channel for 216  insoluble compounds. The relative piston velocity effects of a lipid or marine biopolymer 217  class may well depend on solid versus expanded forms. The sequence of Table 3 entries 218  labeled “2D Phase State” indicates that geographic dependence is expected. All major 219  and trace dissolved gases of the sea are susceptible including those critical to the climate 220  system –carbon dioxide, nitrous oxide, methane, dimethyl sulfide and others (Frew 1997, 221  Leifer and Patro 2002, Tsai and Liu 2003). 222   223  

Page 6: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  6  

A given wind-generated bubble rises steadily back toward the atmosphere, and ultimately 224  it penetrates the overlying ocean-air boundary. There it acquires resident surface material, 225  in the sense of a target cross section. The process can be represented conceptually as the 226  multiplication of the several surfactant monolayers (Oppo et al 1999, Burrows et al 2014). 227  Domed structures form and stretch, leading to a familiar enrichment of marine 228  macromolecules, by many orders of magnitude relative to the total of dissolved organic 229  carbon in seawater (Blanchard 1963 and 1989, Hoffman and Duce 1976, Russell et al 230  2010, Burrows et al 2014). In what follows we sidestep issues such as foam delays, 231  drainage of the aqueous phase from the stressed film or estimation of its thickness 232  (Sellegri et al 2006, Modini et al 2013). This is accomplished by relying primarily on 233  empirical data. 234   235  Table 3 (atmospheric surfactants): 1 Molar/C1/2, (Γmax/1 Molar)(A/V), Πmax 236   237  Carbon enters the marine aerosol system in primary form via bubble breaking, with later 238  additions from secondary level processing through the gas phase. In the main text plus 239  supplemental equations 1 through 15, we attempt to capture the essence of surfactant 240  chemistry along the bubble driven channel. Shifting to the atmosphere, our focus 241  therefore falls naturally upon the primary organic source term. Direct injections become 242  nascent spray particles, and their chemistry remains distinct for an extended period 243  during boundary layer processing (O’Dowd and de Leeuw 2007, Westervelt et al 2012). 244  Meanwhile by many indications, the fractional mass contribution of biomolecules may be 245  comparable to that of sea salt (O’Dowd and de Leeuw 2007, Russell et al 2010, Lapina et 246  al 2011). 247   248  It is readily demonstrated under these circumstances that sea spray organics approach or 249  exceed unit carbon molarity overall. As such, we adopt this value in Table 3 and the main 250  text as a convenient reference concentration level. For example, one can apply a series of 251  traditional rules of thumb to nascent material. 252   253  𝑂:𝐶 ≈ 2;𝜌   ≈ 1.5𝑔 𝑐𝑐 ;  𝑟! ≈ 2𝑟!" ≈ 4𝑟!"# (16) 254   255  (Svenningson et al 2006, O’Dowd and de Leeuw 2007, Westervelt et al 2012). The first 256  entry on line 16 is an average organic to carbon mass relationship for partially oxidized 257  compounds (Turpin and Lim 2003, Russell 2003). Likewise the density provided is a 258  round figure often applied collectively to the remote biogenics (Meskhidze et al 2011). 259  Decreasing radii are defined working from the point of particle generation (ro) then 260  representing water vapor equilibrium at ambient relative humidity and finally the size of 261  the chemistry-determining dry particle core. For equivalent masses of dry organics and 262  sodium chloride diluted to RH 80, the value of C is 5 Molar. Our approximation leaves 263  ample room for reductions to the relative carbon mass or mole fraction. 264   265  Thus the simple half saturation reciprocal 1 (molar carbon)/C1/2 offers a convenient 266  reflection of surface coverage tendency for remote spray. Since KC >> unity in all the 267  Table 3 cases, marine surfactants are initially capable of occupying all positions at the 268  aerosol perimeter. Equations 11 and 15 are turned “inside out”, so to speak, since 269  

Page 7: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  7  

hydrophobic groups now populate the exterior of the microscopic object of interest. 270  Liquid dispersed from a rupturing film/dome into the atmosphere will be under severe 271  mechanical disturbance initially (Spiel 1997a and b, 1998). But fragments soon resolve 272  themselves into spheres and then both vapor and surfactant equilibria are reestablished. 273  Depletion of macromolecular carbon from the interior must of course be accounted 274  (Sorjamaa et al 2004, Petters and Kreidenweis 2013). In the language of our first few 275  supplementary equations ng = 0, n = nl + nsurface = constant. Complete atmospheric models 276  may ultimately deal with this form of mass conservation directly. But here we are content 277  to handle the issue in a zeroth order manner. 278   279  The upper limiting surface reservoir (moles) and round figure bulk reservoir (moles) 280  follow from multiplication through by droplet surface area and volume respectively. A bit 281  of algebra then yields -the ratio of two and three dimensional concentrations followed by 282  a surface to volume term, which brings us at last to spherical geometry. 283   284  𝑛!"#$%&' = Γ!"#𝐴;    𝑛! = 1  Molar  V (17) 285  𝑛!"#$%&' 𝑛! = Γ!"#/1  𝑀𝑜𝑙𝑎𝑟 𝐴 𝑉 (18) 286  𝐴 = 4𝜋𝑟!;𝑉 = 4 3 𝜋𝑟!;  𝐴 𝑉 = 3 𝑟 (19) 287   288  The sea spray injection mode radius can be rounded to the half decade as ro = 0.3 microns 289  (equation 16, Gong 2003, O’Dowd and de Leeuw 2007), so that the quantity 3/r is about 290  107 per meter. One mole per liter of carbon is 1000 moles per cubic meter. Thus 291  nsurface/nlΓmax must be of order 104 m2/mole. The fractions in Table 3 arise from our 292  reliance on microscopically intuitive round values quoted for the excess. For example, 293  one atom per square angstrom is very close to 5/3 x 10-4 moles per square meter. But 294  since the final values for quantity 18 are all small, we can presume that areal saturation is 295  supportable. Barriers to mass transfer and alterations to heterogeneous atmospheric 296  photochemical reaction rates are implicit (Feingold and Chuang 2002, Donaldson and 297  Vaida 2006). Reductions to surface tension follow from the Gibbs equation (8 and 9). Per 298  the two dimensional gas law, they may be expressed as surface pressures (10). 299   300  To assess the effects of Πmax as cited in the main text and Table 3, it becomes necessary 301  to bring curvature into our development for the first time. The physical scale has been 302  shrinking steadily as our carbon moves upward. A swarm of submerged bubbles and the 303  planetary phase boundary may be considered flat collectively. But basic equations 1 304  through 3 must be revisited in light of global aerosolization. The spotlight now falls upon 305  water vapor equilibrium, which determines particle growth versus shrinkage in a rising 306  marine air mass (Pruppacher and Klett 1997, Seinfeld and Pandis 2006). Surfactant 307  terms are subsumed, in fact deleted, in order to focus on the solvent in its two major 308  physical states. The Gibbs phase plane is ignored, but constant temperature and pressure 309  can nonetheless be invoked to achieve standard free energy simplifications. We recognize 310  that the central reaction process has become a bulk or three dimensional phase transition, 311  from liquid to vapor so that -dnl = dng. As always equilibrium is defined by dG =0. The 312  chemical potentials are those of water itself. From 1 through 3 and the definition of µ, 313   314  𝑑𝐺 = 𝜎𝑑𝐴 + 𝜇!𝑑𝑛! + 𝜇!𝑑𝑛!;  𝜎𝑑𝐴/𝑑𝑛! = 𝜇! − 𝜇! (20) 315  

Page 8: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  8  

𝜇! − 𝜇! = 𝜇!"#$%#&%! − 𝜇!"#$%#&%! + 𝑅𝑇𝑙𝑛 𝑝 𝑎 (21) 316  

317  We combine equations 20 and 21 then exponentiate. Standard states drop out in favor of 318  the water vapor pressure since ΔGo concerns products minus reactants and the activity of 319  pure water is one. Otherwise the solvent term reverts to its mole fraction since our 320  mixtures are dilute. So long as we do not enter the realm of molecular clusters, tension σ 321  may be considered fixed -only at the nanometer scale do surface Van der Waals 322  interactions begin to alter with curvature (Seinfeld and Pandis, 2006). Computing the 323  number of moles and drawing on the geometries of 19 together yield 324   325  𝑛! = 𝜌! 𝑀𝑊 𝑉 (22) 326  𝑑𝑛! = 𝜌! 𝑀𝑊 4𝜋𝑟!𝑑𝑟;𝑑𝐴 = 8𝜋𝑟𝑑𝑟;  𝑑𝐴 𝑑𝑛! = 2𝑀𝑊 𝜌! 𝑟 (23) 327  𝑝 𝑣𝑝 = 𝑆𝑅(𝐷) = 𝑎𝑒𝑥𝑝 4𝜎𝑀𝑊 𝑅𝑇𝜌!𝐷 ; 𝑆𝑅(𝐷) = 𝑎𝑒𝑥𝑝(𝐾𝑒𝑙𝑣𝑖𝑛 𝐷) (24) 328  𝐾𝑒𝑙𝑣𝑖𝑛 = 𝐷!"# = 4𝜎𝑀𝑊/𝑅𝑇𝜌! (25) 329   330  where we also take advantage of the opportunity to apply the conversion 2r = D and refer 331  to the equilibrium supersaturation ratio as SR. Both of these adjustments are made for 332  consistency with familiar aerosol science presentations (Petters and Kreidenweis 2007). 333  The Kelvin coefficient is often referred to in the aerosol literature as A (Seinfeld and 334  Pandis 2006, Petters and Kreidenweis 2007, Ghan et al 2011), which is in conflict with 335  our multidisciplinary need to represent total surface area using this symbol. Hence we 336  invoke a reference length instead. 337   338  Equations 24 and 25 highlight the crucial curvature effect on atmospheric droplet vapor 339  pressure. The pure water surface tension is 72 mJ/m2 and so a convenient round figure 340  Dref is 2 nm. If we now round/convert the spray mode injection size r80 = 0.1 to D = 0.2 341  microns (Gong 2003) and linearize the exponential, it is immediately clear that 342  equilibrium could be as high as 1 % supersaturation over nascent material (SR = 1.01). 343  Modulation by the activity term pulls this down by perhaps a factor of three due to the 344  presence of a chemically active core (Seinfeld and Pandis 2006). But with the tendencies 345  1 Molar/C1/2 enforcing surface pressures greater than 30 in some cases (Table 3), SR 346  could be virtually halved over local ecosystems. Pruppacher and Klett (1997) give 0.03 to 347  1 % as the range of maximum supersaturations encountered during updrafts, in marine 348  stratiform to cumulus cloud structures. By this reckoning the macromolecular 349  biogeography of Πmax is central to early cloud formation. 350   351  Table 3 (chemistry in the atmospheric bulk): κ, O:C and ρ 352   353  Activity coefficient reductions in the marine boundary layer are a direct manifestation of 354  Raoult’s Law. They become dominant in the aerosol fine mode, and in the remote 355  oceanic regime aqueous phase chemistry of the macromolecules is necessarily reflected 356  (Meskhidze et al 2011, Westervelt et al 2012). A popular parameterization of the effects 357  allows us to connect with well known physical properties of the organic material (Petters 358  and Kreidenweis 2007). The quantities involved are divergent among themselves and also 359  across the scenarios of Table 3, underscoring again the potential for ecogeographic 360  variation. 361  

Page 9: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  9  

362  The solvent water mole fraction may be modeled as a volume mixing ratio, with 363  appropriate solute weightings accounting for ion dissociation, complexation, 364  reconfiguration, gelling or other critical behaviors that may be observed empirically. 365  Factors involved in the chemical and mass effects are referred to as κ and ε respectively. 366  Since our goal is to extract a solvent vapor pressure, we denote water substance by the 367  special marker w. Individual dissolved species whether organic or otherwise become i 368  and the full complement of solutes is indicated by s. In this final section of the 369  supplement, superscripts are applied to collective components which exist as separate 370  phases only in the conceptual sense, or else under extreme conditions –solutes gather to 371  become a solid core on complete drying. Activity is constructed stepwise assuming that 372  contributing volumes can be summed. Insertion into 24 provides the desired expression. 373  Note geometric cancellations follow once again from 19 (with conversion to D). 374   375  𝑎 = 𝑉! 𝑉! +   𝜅!! 𝑉! = 𝑉! 𝑉! +  𝑉! 𝜅!! 𝜖! = 𝑉!/ 𝑉! +  𝑉!𝜅   (26) 376  𝜅 = 𝜅!𝜖!! ;  𝑉! = 𝑉!! ;  𝜖! = 𝑉! 𝑉! (27) 377  𝑉 = 𝑉! +  𝑉!;  𝐷! = 𝐷!! +  𝐷!! (28)   378  𝑆𝑅(𝐷) = 𝑉! 𝑉! + 𝑉!𝜅 𝑒𝑥𝑝 4𝜎𝑀𝑊 𝑅𝑇𝜌!𝐷 (30) 379  𝑆𝑅(𝐷) = 𝐷! − 𝐷!! 𝐷! − 𝐷!! 1− 𝜅 𝑒𝑥𝑝 𝐷!"! 𝐷 (31) 380   381  The result in 31 is convenient but overly simple, in fact constituting a parameterization of 382  a series of approximations. Nonetheless, it permits us to demonstrate several crucial 383  points with regard to the marine organics. Extensions have been proposed in order to 384  strictly conserve surfactant mass (Sorjamaa et al 2004, Petters and Kreidenweis 2013). 385  Based on conclusions from equation 18, however, this issue may be decoupled for 386  preliminary purposes. 387   388  An effective way to explore the function SR is by plotting versus D, for example fixing 389  Dref = 2 nm while adopting κ values at partial logarithmic intervals. Note that for zero 390  hygroscopicity, aerosol water behaves as though it were pure. Moreover, κi >1 are 391  permitted and in fact documented for certain inert salts. Chemical interactions can 392  produce bumps, kinks and irregularities in the D space (Seinfeld and Pandis 2006). But 393  for purposes here we will analyze the single peak in form 31. Its location can be 394  estimated by 1.) assuming total solute volume is negligible, 2.) application of a binomial 395  approximation to the activity term, 3.) log transformation overall, 4.) linearization, and 396  ultimately 5.) setting the derivative to zero. This procedure produces a relation between 397  the greatest or critical equilibrium supersaturation SRc plus hygroscopicity and the 398  chemical core size (Seinfeld and Pandis 2006, Petters and Kreidenweis 2007, Ghan et al 399  2011). 400   401  𝑙𝑛 𝑆𝑅! =   4𝐷!"#! 27𝜅𝐷!! ! !

;  𝑆! ≈ 1+ 4𝐷!"#! 27𝜅𝐷!! ! ! (32) 402  

403  Equation 32 is acceptable as a learning device over much of the Table 3 range in κ 404  (compare with data in Petters and Kreidenweis 2007). Total hygroscopicity enters 405  directly into the quantity in brackets. Based on our text discussion of the inherent 406  

Page 10: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  10  

uncertainties, the central round figure κ = 0.1 can be adopted for an arbitrary baseline 407  organic. Next the averages from equation 16 are introduced conceptually. Working from 408  the rough Gong (2003) injection radius r80 = 0.1 (D80 = 0.2), a round value core is about 409  Ds = 0.1 microns in size. For the baseline Dref of 2 nm a central Sc 0.3 % follows, which 410  is not unreasonable. But the main issue is sensitivity in any case. The product κ(Ds)3 can 411  be adjusted upward and downward by a factor of 3 to address simultaneously the multiple 412  sources of uncertainty in κ, O:C and density (with the latter two playing into Vs). Sc 413  varies in turn from 0.2 to 0.5 %. Again a significant fraction of the marine updraft 414  spectrum is spanned (Pruppacher and Klett 1997). 415   416  Alert readers from the aerosol community may notice an implicit assumption hidden in 417  the above paragraph -that the organics sometimes possess their own externally mixed 418  state. Dilution with salt weakens the influence for all factors contained in κ(Ds)3. But 419  there will be compensation. For example, high ionic strength likely enhances lipid or 420  protein surface tendencies. Evidence for chemical segregation of the aerosol is strong in 421  any case (Russell et al 2010, Hawkins and Russell 2010, Wex et al 2010). Plus 422  processing toward uniform composition takes time (Westervelt et al 2012), leaving open 423  the possibility for ecosystem scale regimes of biogeochemical influence. Clearly back of 424  the envelope calculations are inadequate to the task at hand, and the problem must be 425  recast in the form of numerical modeling. But this is precisely the point of the entire 426  exercise, and so here we rest our case. 427   428  Supplementary References 429   430  Adamson A 1960 Physical Chemistry of Surfaces (Easton PA: Interscience Publishers) 431  Adamson A and Gast A 1997 Physical Chemistry of Surfaces (New York: Wiley 432  

Interscience) 433  Barger W and Means J 1985 Clues to the structure of marine organic material from the 434  

study of physical properties of surface films Marine and Estuarine Geochemistry ed 435  A. Sigleo and A Hattori (Chelsea MI: Lewis Publishers) 436  

Benner R 2002 Chemical composition and reactivity Biogeochemistry of Dissolved 437  Organic Matter ed D Hansell and C Carlson (New York: Academic) pp 59-90 438  

Birdi K 1989 Lipid and Biopolymer Monolayers at Liquid Interfaces (New York: Plenum 439  Press) 440  

Blanchard D 1963 The electrification of the atmosphere by particles from bubbles in the 441  sea Prog. Oceanogr. 1 73-112 442  

Blanchard D 1989 The ejection of drops from the sea and their enrichment with bacteria 443  and other materials: A review Estuaries 12(3) 127-137 444  

Brzozowska A, Duits, M and Mugele F 2012 Stability of stearic acid monolayers on 445  artificial sea water Coll. Surf. A: Physicochemical and Engineering Aspects 407 38-446  48 447  

Burrows SM et al (2014) A physically-based framework for modeling the organic 448  fractionation of sea spray aerosol from bubble film Langmuir equilibria, Atmos. Chem. 449  Phys. in discussion 450  

Christodoulou A and Rosano H 1968 Effect of pH and nature of monovalent cations on 451  surface isotherms of saturated C16 to C22 soap monolayers Adv. Chem. 84 210-34 452  

Page 11: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  11  

Clift R, Grace J and Weber M 1978 Bubbles, Drops and Particles (New York: Courier 453  Dover) 454  

Cook M and Talbot E 1952 Surface hydrolysis in sodium lauryl sulfate solutions and its 455  effect on surface tension and adsorption at the solid-aqueous solution interface J. 456  Phys. Chem. 56, 412-16. 457  

Cunliffe M, Upstill-Goddard R and Murrell J 2011 Microbiology of aquatic surface 458  microlayers FEMS Microbiol. Rev. 35 233-46 459  

Donaldson D and Vaida V 2006 The influence of organic films at the air-aqueous 460  boundary on atmospheric processes Chem. Rev. 106 1445-61 461  

Feingold G and Chuang P 2002 Analysis of the influence of film-forming compounds on 462  droplet growth: Implications for cloud microphysical processes and climate J. Atmos. 463  Sci. 59 2006-18 464  

Frew N 1997 The role of organic films in air sea gas exchange The Sea Surface and 465  Global Change ed P Liss and R Duce (Cambridge University Press) pp 121-172 466  

Frew N, Goldman J, Dennett M and Johnson A 1990 Impact of phytoplankton generated 467  surfactants on air-sea gas exchange J. Geophys. Res. 95(C3) 3337-52 468  

Ghan S et al 2011 Droplet nucleation: Physically based parameterizations and 469  comparative evaluation J. Adv. Model. Earth Syst. 3 2011NS000074 470  

Gong S 2003 Parameterization of sea-salt aerosol source function for sub- and super-471  micron particles Global Biogeochem. Cycles 19(4) 2003GB002079 472  

Graham D and Phillips M 1979a Proteins at liquid interfaces 1. Kinetics of adsorption 473  and surface denaturation J. Coll. Interface Sci. 70(3) 403-14 474  

Graham D and Phillips M 1979b Proteins at liquid interfaces 2. Adsorption isotherms J. 475  Coll. Interface Sci. 70(3) 415-26 476  

Graham D and Phillips M 1979c Proteins at liquid interfaces 3. Molecular structures of 477  adsorbed films J. Coll. Interface Sci. 70(3) 427-39 478  

Guastalla J 1939 Diluted films of various proteins Compt. Rend. 208 1078-79 479  Hansell D, Carlson C and Schlitzer R 2012 Net removal of major marine dissolved 480  

organic carbon fractions in the subsurface ocean Global Biogeochem. Cycles 26 481  2011GB004069 482  

Hardy J 1982 The sea surface microlayer: Biology, chemistry and anthropogenic 483  enrichment Prog. Oceanogr. 11 307-328 484  

Hawkins L and Russell L 2010 Polysaccharides, proteins and phytoplankton fragments: 485  Four chemically distinct types of marine primary organic aerosol classified by single 486  particle spectromicroscopy Adv. Met. doi:10.1155/2010/612132 487  

Hoffman E and Duce R 1976 Factors influencing the organic carbon content of marine 488  aerosols: A laboratory study J. Geophys. Res. 81(21) 3667-70 489  

Laidler K 1965 Chemical Kinetics (New York: McGraw Hill) 490  Lapina K et al 2011 Investigating organic aerosol loading in the remote marine 491  

environment Atmos. Chem. Phys. Discuss. 11 10973-11006 492  Leifer I and Patro K 2002 The bubble mechanism for methane transport from the shallow 493  

seabed to the surface: A review and sensitivity study Cont. Shelf Res. 22 2409-28 494  Lipitov Y and Sergeeva L 1984 Adsorption of Polymers (New York: John Wiley and 495  

sons) 496  Liss, P 1975 Chemistry of the sea surface microlayer Chemical Oceanography ed J Riley 497  

and G Skirrow (New York: Academic Press) pp 193-244 498  

Page 12: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  12  

Meskhidze N et al 2011 Global distribution and climate forcing of marine organic 499  aerosol: 1. Model improvements and evaluation Atmos. Chem. Phys. 11 11689-705 500  

McGregor M and Barnes G 1978 Equilibrium penetration of monolayers VI: Cholesterol-501  cetrimonium bromide system J. Pharm. Sci. 67(8) 1054-56 502  

Modini R, Russell L, Deane G and Stokes M 2013 Effect of soluble surfactant on bubble 503  persistence and bubble-produced aerosol particles J. Geophys. Res. 118(3) 1388-1400 504  

O’Dowd C and de Leeuw G 2007 Marine aerosol production: A review of current 505  knowledge Phil. Trans. Math. Phys. Eng. Sci. 365 1753-74 506  

Oppo C et al 1999 Surfactant components of marine organic matter as agents for 507  biogeochemical fractionation and pollutant transport via marine aerosols Mar. Chem. 508  63 235-53 509  

Petters M and Kreidenweis S 2007 A single parameter representation of hygroscopic 510  growth and cloud condensation nucleus activity Atmos. Chem. Phys. 7 1961-71 511  

Petters M and Kreidenweis S 2013 A single parameter representation of hygroscopic 512  growth and cloud condensation nucleus activity –Part 3: Including surfactant 513  partitioning Atmos. Chem. Phys. 13 1081-91 514  

Pruppacher H and Klett J 1997 Microphysics of clouds and precipitation (Dordrecht, The 515  Netherlands: Kluwer) 516  

Russell L 2003 Aerosol organic mass to organic carbon ratio measurements Environ. Sci. 517  Technol. 37 2982-87 518  

Russell L, Hawkins L, Frossard A, Quinn P and Bates T 2010 Carbohydrate-like 519  composition of submicron atmospheric particles and their production from ocean 520  bubble bursting PNAS 107(15) 6652-57 521  

Schofield R and Rideal E 1925 The kinetic theory of surface films. Part I. The surfaces of 522  solutions Proc. Roy. Soc. A109 57-77 523  

Seinfeld J 1986 Atmospheric Chemistry and Physics of Air Pollution (Somerset NJ: John 524  Wiley and Sons) 525  

Seinfeld J and Pandis S 2006 Atmospheric Chemistry and Physics: from Air Pollution to 526  Climate Change (Hoboken NJ: John Wiley and Sons) 527  

Sellegri K, O’Dowd C, Yoon Y, Jennings S and de Leeuw G 2006 Surfactants and 528  submicron sea spray generation J. Geophys. Res. 111 2005JD006658 529  

Sorjamaa R, Svenningsson B, Raatikainen T, Henning S, Bilde M and Laaksonen A 2004 530  The role of surfactants in Kohler theory reconsidered Atmos. Chem. Phys. 4 2107-17 531  

Spiel D 1997a A hypothesis concerning the peak in film drop production as a function of 532  bubble size J. Geophys. Res. 102(C1) 1153-61 533  

Spiel D 1997b More on the birth of jet drops from bubbles bursting on seawater surfaces 534  J. Geophys. Res. 102(C3) 5815-21 535  

Spiel D 1998 On the birth of film drops from bubbles bursting on seawater surfaces J. 536  Geophys. Res. 103(C11) 24907-18 537  

Svenningsson B et al 2006 Hygroscopic growth and critical supersaturations for mixed 538  aerosol particles of inorganic and organic compounds of atmospheric relevance Atmos. 539  Chem. Phys. 6 1937-52 540  

Ter Minassian-Saraga L 1956 Recent work on spread monolayers, adsorption and 541  desorption J. Coll. Sci. 11 398-418 542  

Tsai W and Liu K 2003 An assessment of the effect of sea surface surfactant on global 543  atmosphere-ocean CO2 flux J. Geophys. Res. 108(C4) 2000JC000740 544  

Page 13: Theory and equations: supplementary material Table 1 …iopscience.iop.org/1748-9326/9/6/064012/media/ERL479720... · Pointers have been inserted at several appropriate locations

Oceanic macromolecular surfactant estimates    

  13  

Tuckermann R 2007 Surface tension of aqueous solutions of water-soluble organic and 545  inorganic compounds Atmos. Environ. 41 6265-75 546  

Turpin B and Lim H 2001 Species contributions to PM2.5 mass concentrations: 547  Revisiting common assumptions for estimating organic mass Aer. Sci. Tech. 35(1) 548  602-10 549  

Westervelt D, Moore R, Nenes A and Adams P 2012 Effect of primary organic sea spray 550  emissions on cloud condensation nuclei concentrations Atmos. Chem. Phys. 12 89-551  101 552  

Wex H, McFiggans G, Henning S and Stratmann F 2010 Influence of the external mixing 553  state of atmospheric aerosol on derived CCN number concentrations Geophys. Res. 554  Lett. 37 2010GL043337 555  

Woolf D 1997 Bubbles and their role in gas exchange The Sea Surface and Global 556  Change ed P Liss and R Duce (Cambridge University Press) pp 173-205 557