Next Article in Journal
A Satellite Analysis: Comparing Two Medicanes
Previous Article in Journal
Subgrid-Scale Topographic Effects on Radiation for Global Weather Forecast Models
Previous Article in Special Issue
Review of Smog Chamber Experiments for Secondary Organic Aerosol Formation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Relative Humidity on the Rate of New Particle Formation for Different VOCs

by
Austin C. Flueckiger
and
Giuseppe A. Petrucci
*
Department of Chemistry, The University of Vermont, 82 University Place, Burlington, VT 05405, USA
*
Author to whom correspondence should be addressed.
Atmosphere 2024, 15(4), 480; https://doi.org/10.3390/atmos15040480
Submission received: 28 February 2024 / Revised: 3 April 2024 / Accepted: 9 April 2024 / Published: 12 April 2024

Abstract

:
Atmospheric new particle formation (NPF) is an important source of aerosol particles and cloud condensation nuclei, which affect both climate and human health. In pristine environments, oxidation of biogenic volatile organic compounds (VOCs) is a major contributor to NPF. However, the impact of relative humidity (RH) on NPF from these precursors remains poorly understood. Herein, we report on NPF, as inferred from measurements of total particle number density with a particle diameter (dp) > 7 nm, from three VOCs (sabinene, α-terpineol, and myrtenol) subjected to dark ozonolysis. From a series of comparative experiments under humid (60% RH) and dry (~0% RH) conditions and a variety of VOC mixing ratios (ξVOC, parts per billion by volume, ppbv), we show varied behavior in NPF at elevated RH depending on the VOC and ξVOC. In general, RH-dependent enhancement of NPF at an ξVOC between <1 ppbv and 20 ppbv was observed for select VOCs. Our results suggest that gaseous water at particle genesis enhances NPF by promoting the formation of low-volatility organic compound gas-phase products (LVOCs). This is supported by measurements of the rate of NPF for α-pinene-derived SOA, where RH had a greater influence on the initial rate of NPF than did ξVOC and ξO3.

1. Introduction

Understanding the chemical and physical properties of atmospheric aerosols is critical to interpreting their direct and indirect impacts on climate change, such as by scattering and absorbing solar radiation and impacting cloud formation [1,2,3,4,5]. Organic aerosol (OA) accounts for up to 90% of Earth’s total aerosol mass budget [6,7,8], with secondary organic aerosol (SOA) making up 70–90% of OA fine particle mass [9,10]. Therefore, it is of interest to study organic new particle formation (NPF) and SOA mass formation (CSOA), as these processes play a crucial role in the Earth’s atmosphere.
To date, a great deal of effort has been expended to identify sources of organic particles, as well as to better understand the impact of atmospheric parameters, such as temperature, that may influence NPF. In addition to advancing our fundamental knowledge of atmospheric chemical processes, quantitative data on the role of atmospheric parameters on NPF are of interest to improve the predictive accuracy of global climate models [11,12].
Atmospheric NPF is initiated by the formation of molecular clusters, often involving atmospheric ions, sulfuric acid, water, and highly oxidized organic molecules (HOMs) [11,13,14,15,16]; moreover, it has been proposed that the incorporation of HOMs can increase cluster survivability [17,18,19]. Upon cluster formation, a balance exists between losses of these clusters (for example, through processes such as evaporation, condensation, and coagulation) and growth to a critical size that can facilitate atmospheric particle formation via condensation of low-volatility gas-phase products [20]. Significant effort is being expended to understand cluster formation rates and particle growth factors to more accurately quantify and predict atmospheric NPF events [21,22] (and references therein). However, the ensuing discussion adopts a more practical view of NPF and subsequent particle growth as the emergence of new aerosol particles into the lower end of the measured particle size spectrum, followed by the growth of the particles [21,23].
It is generally accepted that the critical cluster size that favors survivability and growth is sub-3 nm [13,21]. Simplistically, once a critical cluster size is attained, the rate at which new particles are formed and grow to measurable sizes (Jap) depends on the relative rates of production (P) of clusters and condensable species compared to vapor and cluster losses (L (i.e., sinks)) [23]). Hence, without fully understanding the underlying mechanisms of cluster formation and survivability, Jap can be expressed as a competition between these two factors, as seen in Equation (1):
J a p P L
By extension, any parameter that impacts either the P or the L process (or both) must, by necessity, also impact Jap and NPF. For example, recent studies have shown a clear negative correlation between temperature and NPF, suggested to be due to the increased rate of production of stable clusters and lower equilibrium vapor pressures of condensable species [24,25,26].
Gaseous water is the third most abundant atmospheric gas; however, the mechanistic role it plays in organic NPF is not well understood. Recent reports on the effect of relative humidity (RH) on NPF have been contradictory, with a few studies reporting a positive correlation between NPF and increased humidity [27,28,29,30,31], while most report a negative correlation [29,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46]. Recent work from our laboratory demonstrated both enhancement and attenuation in NPF for the dark ozonolysis of α-pinene depending on the VOC mixing ratio (ξVOC, parts per billion by volume, ppbv) [29].
Herein, we took an empirical approach toward observing the impact of RH on SOA and NPF behavior from the ozonolysis of a variety of volatile organic compound (VOC) precursors. Specifically, we report on the effect of RH on organic NPF for three VOCs (sabinene, α-terpineol, and myrtenol) that were subjected to dark ozonolysis at various ξVOCs and ozone mixing ratios (ξO3), and compare against previous studies that used α- and β-pinene [29] and cis-3-hexenyl acetate (CHA) [46]. Sabinene, α-terpineol, and myrtenol were selected for this work, as they are structural derivatives of α-pinene (RH has been shown to affect the NPF of α-pinene-derived SOA [29]), but they have varying solubilities. Of the three VOCs studied, sabinene and α-terpineol displayed enhancements in NPF under humid conditions (60% RH) relative to dry conditions (~0% RH), specifically at ξVOCs of less than a few ppbv. However, myrtenol showed an attenuation in NPF under humid conditions at all ξVOCs. These results are in agreement with our recent observations from the dark ozonolysis of α- and β-pinene [29] and cis-3-hexenyl acetate (CHA) [46]. Our results suggest a variable correlation of NPF to increased RH at particle genesis, with a dependence on the VOC undergoing ozonolysis. In addition, the impact of ξVOC, ξO3, and RH on the Jap was examined using α-pinene as a model for VOCs that have shown NPF enhancements. NPF rate data, coupled with the observation that the highly water-soluble α-terpineol showed the greatest enhancement in NPF under humid conditions, suggests gaseous water plays a role in NPF and SOA formation through changing chemical mechanisms and/or reaction rates.

2. Materials and Methods

2.1. Reagents

The following reagents were used for this work without further purification: α-pinene (98%, CAS: 7785-70-8, Alfa Aesar, Haverhill, MA, USA), β-pinene (99%, CAS: 19902-08-0, Sigma Aldrich, St. Louis, MO, USA), sabinene (75%, CAS: 3387-41-5, Sigma Aldrich, St. Louis, MO, USA), myrtenol (97.1%, CMX: 34687, Chem-Impex International Inc., Wood Dale, IL, USA), CHA (>97.0%, CAS: 3681-71-8, TCI, Portland, OR, USA), and α-terpineol (97+%, CAS: 98-55-5, Acros Organics, Waltham, MA, USA).

2.2. Instrumentation and Experimental Conditions

Particle number concentration (N, # cm−3), SOA mass (CSOA, µg m−3), and particle geometric mean diameter (GMD, nm) were measured or derived continuously using a scanning mobility particle sizer (SMPS 3082, TSI Inc., Shoreview, MN, USA). The SMPS utilized a long DMA and a 3756 UCPC operating at 0.3 and 3 L/min for the aerosol and sheath flow rates, respectively, resulting in a particle mobility diameter range of 17–583 nm. Rates of particle formation (Jap) were determined from measurements made with an electrical low-pressure impactor (ELPI+, Dekati Technologies Ltd., Kangasala, Finland) consisting of 15 impaction stages ranging from <6 nm to 15 µm. Ozone (O3) was produced using a commercial generator (Ozone Technologies LLC, Model 1KNT, Jersey City, NJ, USA) by passing dry, particle-free air through a corona discharge; ozone concentration was continuously monitored using a commercial unit (Serinus 10, American Ecotech, Warren, RI, USA). RH and temperature inside the chamber were monitored using a Vaisala HUMICAP® HMT130 dual sensor (Vaisala Inc., Vantaa, Finland). The resolution of the RH measurement was ±0.1%. All experiments were run in batch mode and performed under ambient temperature (22 ± 1 °C) and pressure (992 ± 5 mbar) in the University of Vermont Environmental Chamber (UVMEC, Figure 1), an 8 m3 Teflon chamber equipped with multiple reagent and sampling ports. The VOC injection port was run through a magnetically coupled fan to facilitate VOC/O3 mixing. For all experiments (unless otherwise stated), ξO3 = 500 (±30) ppbv and the RH was either dry (~0%) or humid (60%).

2.3. Data Analysis and Terminology

An enhancement factor (ΔNmax, Equation (2)) was used to quantify changes in measured NPF induced by RH:
Δ N m a x = ( N m a x , 60 R H N m a x , 0 R H ) N m a x , 0 R H
here, Nmax,60RH and Nmax,0RH are the maximum particle number concentrations measured over the course of a humid and dry experiment, respectively. Both Nmax,60RH and Nmax,0RH were measured under nominally identical ξVOC and ξO3, as ΔNmax is a comparative measurement. Each ΔNmax was determined from the average Nmax,60RH and Nmax,0RH of two runs at each ξVOC; typical relative standard deviation for each metric was less than 10% [46].
The amplification factor (ΔE), used to describe the sensitivity of NPF to changes in each parameter, was determined using Equation (3):
Δ E = ( N m a x , P N m a x , 0 ) N m a x , 0
where Nmax,P is the maximum particle number concentration measured at the indicated ξVOC, ξO3, or RH, and Nmax,0 is the maximum particle number concentration measured at the initial point in each respective (ξVOC, ξO3, or RH) series. For example, if Nmax = 1.0 × 103, 6.0 × 103, and 1.5 × 104 particles cm−3 for a series of three ξVOCs (e.g., ξVOC = 10, 20, and 30 ppbv) run at constant ξO3 and RH (500 ppbv and <1%, respectively), then the corresponding amplification factor for each ξVOC is ΔE = 0, 5, and 14. Therefore, it is worth noting that ΔE = 0 at the initial point in each series.

2.4. Methodology

In this work, each VOC was subjected independently to dark ozonolysis under dry and humid conditions, from which ΔNmax was determined using Equation (2). For a typical experiment, the UVMEC was sealed while slightly over pressure (+2 mbar) to maintain a consistent initial chamber volume. The chamber fan was then turned on and O3 was injected for a pre-determined time pulse (typically 100–120 s) to reach ξO3 = 500 ± 30 ppbv. After an equilibration period of 10–15 min, the particle concentration was measured with the SMPS to ensure a clean background spectrum (<10 particles cm−3). An optimized method of VOC injection [46] was used to improve the reproducibility of NPF and SOA generation. Briefly, the liquid aliquot of VOC was injected into a heated three-neck bulb-flask under zero-flow conditions. After 45 s, airflow was diverted through the flask to rapidly transfer a bolus of evaporated VOC to the UVMEC via a calibrated split flow valve. To reproducibly attain the low mixing ratios used in this work, the output flow from the flask passed through a split valve that allowed only a small portion (~25%) of the vaporized VOC to enter the UVMEC. The volume of liquid VOC needed to attain the desired ξVOC within the UVMEC was calculated prior to VOC injection based on the measured split ratio for each experiment. Regardless of the split ratio, the flow to the UVMEC was kept constant at 2.000 +/− 0.015 L min−1. Aerosol size distributions were monitored continuously for 1 h after injection of the VOC using the SMPS and ELPI+. Although we did not correct for wall losses explicitly, it was assumed that wall losses were nominally constant throughout all experiments.

3. Results and Discussion

3.1. Enhancement Factor of NPF (ΔNmax) for Sabinene, α-Terpineol, and Myrtenol as a Function of ξVOC

Of the three VOCs examined, sabinene and α-terpineol displayed enhancements in NPF (i.e., positive ΔNmax) under humid conditions with decreasing ξVOC; this is in good agreement with a recent study focused on α- and β-pinene dark ozonolysis [29]. Contrary to this behavior, myrtenol exhibited an attenuation in NPF under humid conditions at all ξVOCs studied (Figure 2c), analogous to the behavior of CHA in similar experiments [46]. Figure 2 shows the Nmax produced for dry (tan, solid) and humid (green, hashed) experiments for the three VOCs as a function of ξVOC. Generally, as ξVOC decreased, the maximum absolute number of particles produced (regardless of RH) also decreased. However, for sabinene and α-terpineol, the particles produced under humid conditions comparatively increased relative to dry conditions as ξVOC decreased.
Table 1 summarizes the ΔNmax (determined from Equation (2)) for each VOC with respect to ξVOC (shown graphically in Figure 3), along with previously reported enhancement data for α-pinene, β-pinene, and CHA. It should be noted that for cases of ξVOC where no NPF was observed (for example, α-terpineol < 1 ppbv), it was not possible to calculate ΔNmax, even if NPF was observed under humid conditions. Here, it becomes evident that the impact of RH on NPF from the ozonolysis of biogenic VOCs is variable and dependent on the molecular structure of the precursor. Although there was an attenuation in NPF and SOA formation at higher ξVOCs for all VOCs, at lower ξVOCs, only CHA and myrtenol OA continued to exhibit an attenuation in NPF and SOA mass in the presence of elevated RH at particle genesis. This latter observation is in accordance with other literature reports when working at higher ξVOCs (≥50 ppbv) [29,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46]. However, we demonstrate that even for those VOCs that suffer an attenuation in NPF and SOA mass formation under humid conditions at high ξVOCs, at lower, atmospherically relevant ξVOCs, there is an inversion in behavior, and an enhancement in NPF and SOA is measured in the presence of water at particle genesis. Across all VOCs studied here, the magnitude and sign of the RH-induced effect was dependent upon the specific VOC undergoing ozonolysis.

3.2. Addressing Solubility of VOCs

It is possible that the shutdown of NPF for some VOCs is due to their solubility. As discussed in a previous report [46], Teflon walls can develop a significant water film under humid conditions. This film could enhance vapor phase losses to the chamber walls of water-soluble VOCs, thereby causing a decrease in NPF. However, in accordance with the results from this previous work, the current results support the notion that the aqueous solubility of the individual VOCs does not influence the attenuation of SOA formation. As seen in Figure 4a, α-pinene, β-pinene, and sabinene, which are nearly insoluble in water (Table 2), produced an enhancement in NPF under humid conditions at low ξVOCs. CHA and myrtenol are both soluble in water (Table 2), and the ozonolysis of these VOCs at high RH resulted in an attenuation of NPF at all ξVOCs. However, α-terpineol, which is more soluble than CHA and myrtenol (by an order of magnitude), showed the greatest enhancement in NPF under humid conditions of all the VOCs studied. This would indicate that the attenuation of NPF and SOA formation is not due to loss of product to the chamber walls under humid conditions (i.e., a solubility issue) but rather that the mechanism for which gaseous water affects the formation of SOA is unique to individual VOCs.

3.3. Effect of RH on the Rate of NPF for α-Pinene-Derived SOA

3.3.1. Effect of ξVOC, ξO3, and RH on ΔNmax

Three of the controllable variables that affect the formation of SOA at particle genesis in our batch-mode atmospheric chamber experiments were ξVOC, ξO3, and RH. It was determined that all three variables had an influence on ΔNmax (Figure 5). Under constant ξO3 (500 ppbv), there existed a non-linear increase in ΔNmax as ξVOC decreased (Figure 5a). By increasing ξO3 while keeping a fixed ξVOC (10 ppbv), ΔNmax generally increased (Figure 5b), as one would expect. Similarly, Figure 5c shows an increase in ΔNmax with increasing “humid” RH (ξVOC and ξO3 fixed at 10 and 500 ppbv, respectively) compared to ~0% RH. Although the greatest enhancement of ΔNmax was influenced by a decrease in ξVOC, it is evident that all three chamber variables had an influence on ΔNmax at low ξVOCs. It is important to note that the relative importance of each parameter may depend on the absolute value of the parameter that remains fixed. For example, the ΔNmax corresponding to ξVOC may have differed if the fixed ξO3 = 10 ppbv rather than 500 ppbv. For the present study, each fixed parameter was chosen to permit comparison with other work in our laboratory [29,46,47] and in the literature.

3.3.2. Effect of ξVOC, ξO3, and RH on ΔE

The sensitivity of NPF from the dark ozonolysis of α-pinene to each chamber variable was quantified by way of the amplification factor (ΔE, comparative measure of the maximum particle number concentration produced at the given ξVOCO3/RH relative to the initial point in each respective series). ΔE was derived from Equation (3) and assessed for each chamber variable (Figure 6). The y-axis is scaled to provide a direct, visual comparison of the enhancement between the three variables. Figure 6a shows that an increase in ξVOC with static ξO3 and RH (500 ppbv and ~0%) led to an increase in ΔE. Similarly, Figure 6c shows that an increase in RH with static ξVOC and ξO3 (10 ppbv and 500 ppbv) also yielded a general increase in ΔE. However, somewhat surprisingly, this is not the case for an increase in ξO3 (Figure 6b), which yielded little to no change in ΔE (static ξVOC of 10 ppbv and RH of ~0%), suggesting that regardless of the absolute value of ξO3, the system’s sensitivity to changes in the parameter remains constant.

3.3.3. Effect of ξVOC, ξO3, and RH on Jap

To shed light on the effect of RH on NPF for the VOCs that exhibited an enhancement under humid conditions, a systematic study was conducted by which the effect of varying chamber conditions on the apparent rate of particle formation (Jap) was examined. For this work, α-pinene-derived SOA was used as an example for the VOCs that exhibited an enhancement in NPF.
The general trends in normalized Jap (Figure 7) match comparatively with the general trends for ΔE in each series (Figure 6), where increases in ξVOC and RH enhanced the apparent rate of particle formation but ξO3 did not. Although the absolute increase in Jap with respect to ξVOC was greater than that with respect to RH (Figure 7a compared to Figure 7c), when plotting Jap vs. ΔE (Figure 8), the slope of the log–log plot for RH-dependent Jap was approximately 16 times greater than the ξVOC-dependent Jap. This indicates that at low ξVOC of α-pinene, at particle genesis, changes in RH have a much greater impact on the apparent rate at which the OA particles form compared to the ξVOC and ξO3. This would indicate that RH-dependent rates of NPF are likely a major driving force of the eight-fold enhancement in NPF for α-pinene-derived SOA (seen here: Snyder et al. [29]).
In a study conducted by Chu et al. [48], the authors discuss the increased deviation in experimental aerosol formation from their predicted models corresponding to increased RH. The possible explanation posed is that the gaseous water would be taken up by particles, leading to greater surface area and subsequently growing the condensation sink for gas-phase products to partition onto. Although plausible at higher ξVOCs, where we also see an attenuation in NPF with elevated RH for all VOCs, that explanation does not fit our observations at low ξVOCs. At low ξVOCs, for certain VOCs, it appears as though the presence of gaseous water (in excess compared to the concentration of the VOC) led to a promotion of OH reactivity prior to ozonolysis of the VOC [49,50], which seemingly promotes low-volatility organic compound (LVOC) gas-phase product formation and subsequent nucleating species that yield NPF events at a greater rate than when ξVOCs are higher. This is supported by the observation (Figure 8) that ΔE showed a 16× greater dependence on RH than ξVOC.
Regarding pristine environments, pure organic NPF has also been observed [51,52,53,54,55,56], and the theoretical underpinning of organic cluster formation and nucleation has recently been reviewed [19]. Here, it is proposed that pure organic clusters are formed and perhaps stabilized by water molecules or dimers in direct relation to the vapor pressure of the oxidized products and their gas-phase concentrations. If the instantaneous gas-phase concentration of an oxidation product is greater than its equilibrium vapor pressure (i.e., the system is supersaturated in that product), then the probability of cluster formation and NPF increases (see, for example, [48,55,57,58,59]). This supposition is supported by recent laboratory studies that showed that decreases in system temperature (i.e., decreases in equilibrium vapor pressure) resulted in significant enhancements in NPF [24,25,26]. This enhancement was observed despite decreases in the formation of HOMs at lower temperatures [60].
Elevated RH is likely to increase L (Equation (1)) through processes such as enhanced coagulation scavenging of newly formed clusters and condensable gases [61], as well as increased coagulation and scavenging rates of ultrafine particles [42,62]. For example, water vapor uptake to particles could reduce organic (equilibrium) partial pressures according to Raoult’s law. Particle deliquescence would significantly dilute the organics in the particles, resulting in a reduction in organic partial pressures according to Raoult’s law [63,64,65,66]. In addition, an aqueous phase could attract water-soluble organics according to Henry’s law. Lastly, humidity on chamber walls may have an outsized impact on losses of ultrafine particles, which generally suffer greater wall loss rates than larger particles [67]. Therefore, as wall losses could increase at higher RH, the results reported herein could be interpreted as lower limits of the enhancements in NPF as a function of RH.
By all accounts, increases in particle liquid water (PLW) enhance the partitioning of semi- and low-volatility gas-phase products (S- and LVOCs) to the organic particulate (due to a lowering of the average molecular weight of particle constituents) [68,69,70,71,72]. Furthermore, PLW has been shown to increase particle diffusivity, enhancing condensation rates to organic particles [65,71,73,74,75]. Again, both of these effects would result in an increased sink (i.e., L, Equation (1)) at higher RH. The main L for newly formed particles in the atmosphere is coagulation onto an existing particle population [76], whereas in chamber experiments, the main L stems from particle losses to chamber walls, which are normally much more important than coagulation [77] because the existing particle population is usually absent or small. It cannot be discounted that elevated RH could lead to decreased vapor-phase wall losses, thereby increasing the concentration of LVOCs in the gas phase (CLVOC considered analogous to the yield of LVOCs, φLVOC, used by Chu et al. [71]), which could favor NPF. There is only one case, however, of which the authors are aware where increased relative humidity resulted in very modest mitigation of gas-phase wall losses [61]. This singular case studied the loss of oxygenated vapors resulting from the reaction of OH with toluene, a very different chemical system from the current study.
Therefore, RH is likely playing a role by increasing P (Equation (1)) by any number of processes, such as modified oxidized product distributions, enhanced formation of nucleating species, enhanced rate of cluster formation, and/or enhanced survivability (i.e., stabilization) of clusters [31]. Regardless of the precise mechanism(s) by which NPF is impacted by the presence of water, therefore, it appears likely that an increase in water concentration must, by a gas-phase reaction, produce more LVOCs or the rates of production of LVOCs must be greater, effectively outcompeting vapor sinks such as condensation and partitioning in the presence of seed particles or, perhaps, coagulation and wall loss effects (both gas and particle) common in chamber experiments (see Charan et al. [78] for an excellent review on the topic).

4. Conclusions

Overall, the correlation between RH and organic NPF is variable, with a significant dependence on the specific VOC undergoing ozonolysis. In this work, three VOCs (sabinene, α-terpineol, and myrtenol) were subjected to dark ozonolysis under dry and humid conditions (~0 and 60% RH, respectively) at a variety of ξVOCs. When incorporating data from our previous work [29,46], four VOCs (α-pinene, β-pinene, sabinene, and α-terpineol) displayed enhancements in NPF under humid conditions as the ξVOC decreased; however, myrtenol and CHA showed an attenuation in NPF under humid conditions for all ξVOCs. For the VOCs that had NPF enhancements, gaseous water likely played a crucial role in the chemical formation of low-volatility gas-phase products at low ξVOCs. Therefore, these low-volatility products would have spawned new and more frequent NPF events compared to those at high ξVOCs and/or dry conditions. In addition, we note that the observed behavior is not solely due to a VOC’s solubility; this is supported by our observation where α-terpineol produced the greatest ΔNmax despite its solubility being an order of magnitude greater than that of CHA and myrtenol.
Furthermore, as a proof-of-concept study, α-pinene was used as a model to systematically probe the effect of each chamber variable (ξVOC, ξO3, and RH) on the observed enhancement in NPF. Here, it was determined that the ΔNmax and ΔE were dependent on all three chamber variables (ξVOC, ξO3, and RH), while Jap was independent of ξO3. However, of the two dependences (ξVOC and RH), when compared to their respective amplification factors, the sensitivity of NPF was influenced more by RH than by ξVOC (approximately 16×). Therefore, for the VOCs that exhibit NPF enhancements at low ξVOCs, the presence of gaseous water at particle genesis during ozonolysis is shown to be the driving force behind the production of enhanced NPF events.
The role of humidity in atmospheric organic NPF is complex and dependent on the specific SOA precursor under consideration. Furthermore, while the current study focused primarily on two RH levels (~0% reflective of most laboratory studies of SOA and 60% reflective of the average global RH), the preliminary results from α-pinene clearly point to the need to study systematically these effects at intermediate RHs reflective of different environments. It may be possible that upon more extensive experimentation with a broader range of VOCs and across a secondary dimension of RH for each VOC, general parameters may be derived to facilitate incorporation of these effects into climate models to improve outcomes.

Author Contributions

A.C.F. and G.A.P. contributed equally to the ideation of the experiments. A.C.F. carried out all experiments. G.A.P. supervised all laboratory activities. All authors contributed to the preparation of this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This material is based upon work supported by the National Science Foundation under Grant No. CHE-1709751.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Forster, P.; Storelvmo, T.; Armour, K.; Collins, W.; Dufresne, J.-L.; Frame, D.; Lunt, D.J.; Mauritsen, T.; Palmer, M.D.; Watanabe, M.; et al. Climate Change 2021—The Physical Science Basis: Working Group i, the Earth’s Energy Budget, Climate Feedbacks and Climate Sensitivity; IPCC: Geneva, Switzerland, 2023; pp. 923–1054. [Google Scholar]
  2. Zhang, C.Q.; Guo, Y.D.; Shen, H.R.; Luo, H.; Pullinen, I.; Schmitt, S.H.; Wang, M.J.; Fuchs, H.; Kiendler-Scharr, A.; Wahner, A.; et al. Contrasting Influence of Nitrogen Oxides on the Cloud Condensation Nuclei Activity of Monoterpene-Derived Secondary Organic Aerosol in Daytime and Nighttime Oxidation. Geophys. Res. Lett. 2023, 50, e2022GL102110. [Google Scholar] [CrossRef]
  3. Jo, D.S.; Nault, B.A.; Tilmes, S.; Gettelman, A.; Mccluskey, C.S.; Hodzic, A.; Henze, D.K.; Nawaz, M.O.; Fung, K.M.; Jimenez, J.L. Global Health and Climate Effects of Organic Aerosols from Different Sources. Environ. Sci. Technol. 2023, 57, 13793–13807. [Google Scholar] [CrossRef] [PubMed]
  4. Li, J.; Carlson, B.E.; Yung, Y.L.; Lv, D.R.; Hansen, J.; Penner, J.E.; Liao, H.; Ramaswamy, V.; Kahn, R.A.; Zhang, P.; et al. Scattering and absorbing aerosols in the climate system. Nat. Rev. Earth Environ. 2022, 3, 363–379. [Google Scholar] [CrossRef]
  5. Shrivastava, M.; Cappa, C.D.; Fan, J.; Goldstein, A.H.; Guenther, A.B.; Jimenez, J.L.; Kuang, C.; Laskin, A.; Martin, S.T.; Ng, N.L.; et al. Recent advances in understanding secondary organic aerosol: Implications for global climate forcing. Rev. Geophys. 2017, 55, 509–559. [Google Scholar] [CrossRef]
  6. Ehn, M.; Thornton, J.A.; Kleist, E.; Sipila, M.; Junninen, H.; Pullinen, I.; Springer, M.; Rubach, F.; Tillmann, R.; Lee, B.; et al. A large source of low-volatility secondary organic aerosol. Nature 2014, 506, 476–479. [Google Scholar] [CrossRef] [PubMed]
  7. Hallquist, M.; Wenger, J.C.; Baltensperger, U.; Rudich, Y.; Simpson, D.; Claeys, M.; Dommen, J.; Donahue, N.M.; George, C.; Goldstein, A.H.; et al. The formation, properties and impact of secondary organic aerosol: Current and emerging issues. Atmos. Chem. Phys. 2009, 9, 5155–5236. [Google Scholar] [CrossRef]
  8. Jimenez, J.L.; Canagaratna, M.R.; Donahue, N.M.; Prevot, A.S.H.; Zhang, Q.; Kroll, J.H.; DeCarlo, P.F.; Allan, J.D.; Coe, H.; Ng, N.L.; et al. Evolution of Organic Aerosols in the Atmosphere. Science 2009, 326, 1525–1529. [Google Scholar] [CrossRef]
  9. Kanakidou, M.; Seinfeld, J.H.; Pandis, S.N.; Barnes, I.; Dentener, F.J.; Facchini, M.C.; Van Dingenen, R.; Ervens, B.; Nenes, A.; Nielsen, C.J.; et al. Organic aerosol and global climate modelling: A review. Atmos. Chem. Phys. 2005, 5, 1053–1123. [Google Scholar] [CrossRef]
  10. Shrivastava, M.; Easter, R.C.; Liu, X.H.; Zelenyuk, A.; Singh, B.; Zhang, K.; Ma, P.L.; Chand, D.; Ghan, S.; Jimenez, J.L.; et al. Global transformation and fate of SOA: Implications of low-volatility SOA and gas-phase fragmentation reactions. J. Geophys. Res. Atmos. 2015, 120, 4169–4195. [Google Scholar] [CrossRef]
  11. Olenius, T.; Bergström, R.; Kubecka, J.; Myllys, N.; Elm, J. Reducing chemical complexity in representation of new-particle formation: Evaluation of simplification approaches. Environ. Sci. Atmos. 2023, 3, 552–567. [Google Scholar] [CrossRef]
  12. Bender, F.A.M.; Frey, L.; McCoy, D.T.; Grosvenor, D.P.; Mohrmann, J.K. Assessment of aerosol-cloud-radiation correlations in satellite observations, climate models and reanalysis. Clim. Dyn. 2019, 52, 4371–4392. [Google Scholar] [CrossRef]
  13. Kulmala, M.; Petaja, T.; Ehn, M.; Thornton, J.; Sipila, M.; Worsnop, D.R.; Kerminen, V.M. Chemistry of Atmospheric Nucleation: On the Recent Advances on Precursor Characterization and Atmospheric Cluster Composition in Connection with Atmospheric New Particle Formation. Annu. Rev. Phys. Chem. 2014, 65, 21–37. [Google Scholar] [CrossRef] [PubMed]
  14. Stolzenburg, D.; Cai, R.L.; Blichner, S.M.; Kontkanen, J.; Zhou, P.T.; Makkonen, R.; Kerminen, V.M.; Kulmala, M.; Riipinen, I.; Kangasluoma, J. Atmospheric nanoparticle growth. Rev. Mod. Phys. 2023, 95, 045002. [Google Scholar] [CrossRef]
  15. Kirkby, J.; Amorim, A.; Baltensperger, U.; Carslaw, K.S.; Christoudias, T.; Curtius, J.; Donahue, N.M.; El Haddad, I.; Flagan, R.C.; Gordon, H.; et al. Atmospheric new particle formation from the CERN CLOUD experiment. Nat. Geosci. 2023, 16, 948–957. [Google Scholar] [CrossRef]
  16. Zhang, R.Y.; Khalizov, A.; Wang, L.; Hu, M.; Xu, W. Nucleation and Growth of Nanoparticles in the Atmosphere. Chem. Rev. 2012, 112, 1957–2011. [Google Scholar] [CrossRef]
  17. Riccobono, F.; Schobesberger, S.; Scott, C.E.; Dommen, J.; Ortega, I.K.; Rondo, L.; Almeida, J.; Amorim, A.; Bianchi, F.; Breitenlechner, M.; et al. Oxidation Products of Biogenic Emissions Contribute to Nucleation of Atmospheric Particles. Science 2014, 344, 717–721. [Google Scholar] [CrossRef] [PubMed]
  18. Kupc, A.; Williamson, C.J.; Hodshire, A.L.; Kazil, J.; Ray, E.; Bui, T.P.; Dollner, M.; Froyd, K.D.; McKain, K.; Rollins, A.; et al. The potential role of organics in new particle formation and initial growth in the remote tropical upper troposphere. Atmos. Chem. Phys. 2020, 20, 15037–15060. [Google Scholar] [CrossRef]
  19. Elm, J.; Ayoubi, D.; Engsvang, M.; Jensen, A.B.; Knattrup, Y.; Kubecka, J.; Bready, C.J.; Fowler, V.R.; Harold, S.E.; Longsworth, O.M.; et al. Quantum chemical modeling of organic enhanced atmospheric nucleation: A critical review. Wires Comput. Mol. Sci. 2023, 13, 1662. [Google Scholar] [CrossRef]
  20. Li, C.X.; Zhao, Y.; Li, Z.Y.; Liu, L.; Zhang, X.H.; Zheng, J.; Kerminen, V.M.; Kulmala, M.; Jiang, J.K.; Cai, R.L.; et al. The dependence of new particle formation rates on the interaction between cluster growth, evaporation, and condensation sink. Environ. Sci. Atmos. 2023, 3, 168–181. [Google Scholar] [CrossRef]
  21. Dada, L.; Lehtipalo, K.; Kontkanen, J.; Nieminen, T.; Baalbaki, R.; Ahonen, L.; Duplissy, J.; Yan, C.; Chu, B.W.; Petaja, T.; et al. Formation and growth of sub-3-nm aerosol particles in experimental chambers. Nat. Protoc. 2020, 15, 1013–1040. [Google Scholar] [CrossRef]
  22. Li, C.X.; Signorell, R. Understanding vapor nucleation on the molecular level: A review. J. Aerosol Sci. 2021, 153, 105676. [Google Scholar] [CrossRef]
  23. Kerminen, V.M.; Chen, X.M.; Vakkari, V.; Petaja, T.; Kulmala, M.; Bianchi, F. Atmospheric new particle formation and growth: Review of field observations. Environ. Res. Lett. 2018, 13, 103003. [Google Scholar] [CrossRef]
  24. Deng, Y.G.; Inomata, S.; Sato, K.; Ramasamy, S.; Morino, Y.; Enami, S.; Tanimoto, H. Temperature and acidity dependence of secondary organic aerosol formation from alpha-pinene ozonolysis with a compact chamber system. Atmos. Chem. Phys. 2021, 21, 5983–6003. [Google Scholar] [CrossRef]
  25. Heinritzi, M.; Dada, L.; Simon, M.; Stolzenburg, D.; Wagner, A.C.; Fischer, L.; Ahonen, L.R.; Amanatidis, S.; Baalbaki, R.; Baccarini, A.; et al. Molecular understanding of the suppression of new-particle formation by isoprene. Atmos. Chem. Phys. 2020, 20, 11809–11821. [Google Scholar] [CrossRef]
  26. Kristensen, K.; Jensen, L.N.; Quelever, L.L.J.; Christiansen, S.; Rosati, B.; Elm, J.; Teiwes, R.; Pedersen, H.B.; Glasius, M.; Ehn, M.; et al. The Aarhus Chamber Campaign on Highly Oxygenated Organic Molecules and Aerosols (ACCHA): Particle formation, organic acids, and dimer esters from alpha-pinene ozonolysis at different temperatures. Atmos. Chem. Phys. 2020, 20, 12549–12567. [Google Scholar] [CrossRef]
  27. Jonsson, A.M.; Hallquist, M.; Ljungstrom, E. Impact of humidity on the ozone initiated oxidation of limonene, Delta(3)-carene, and alpha-pinene. Environ. Sci. Technol. 2006, 40, 188–194. [Google Scholar] [CrossRef] [PubMed]
  28. Yu, K.P.; Lin, C.C.; Yang, S.C.; Zhao, P. Enhancement effect of relative humidity on the formation and regional respiratory deposition of secondary organic aerosol. J. Hazard. Mater. 2011, 191, 94–102. [Google Scholar] [CrossRef] [PubMed]
  29. Snyder, C.N.; Flueckiger, A.C.; Petrucci, G.A. Relative Humidity Impact on Organic New Particle Formation from Ozonolysis of α- and β-Pinene at Atmospherically Relevant Mixing Ratios. Atmosphere 2023, 14, 173. [Google Scholar] [CrossRef]
  30. Huang, Y.; Ho, K.F.; Ho, S.S.H.; Lee, S.C.; Yau, P.S.; Cheng, Y. Physical parameters effect on ozone-initiated formation of indoor secondary organic aerosols with emissions from cleaning products. J. Hazard. Mater. 2011, 192, 1787–1794. [Google Scholar] [CrossRef]
  31. Jonsson, Å.M.; Hallquist, M.; Ljungström, E. The effect of temperature and water on secondary organic aerosol formation from ozonolysis of limonene, Δ3-carene and α-pinene. Atmos. Chem. Phys. 2008, 8, 6541–6549. [Google Scholar] [CrossRef]
  32. Bousiotis, D.; Brean, J.; Pope, F.D.; Dall’Osto, M.; Querol, X.; Alastuey, A.; Perez, N.; Petaja, T.; Massling, A.; Nojgaard, J.K.; et al. The effect of meteorological conditions and atmospheric composition in the occurrence and development of new particle formation (NPF) events in Europe. Atmos. Chem. Phys. 2021, 21, 3345–3370. [Google Scholar] [CrossRef]
  33. Zhao, Z.X.; Le, C.; Xu, Q.; Peng, W.H.; Jiang, H.H.; Lin, Y.H.; Cocker, D.R.; Zhang, H.F. Compositional Evolution of Secondary Organic Aerosol as Temperature and Relative Humidity Cycle in Atmospherically Relevant Ranges. ACS Earth Space Chem. 2019, 3, 2549–2558. [Google Scholar] [CrossRef]
  34. Liang, L.L.; Engling, G.; Cheng, Y.; Zhang, X.Y.; Sun, J.Y.; Xu, W.Y.; Liu, C.; Zhang, G.; Xu, H.; Liu, X.Y.; et al. Influence of High Relative Humidity on Secondary Organic Carbon: Observations at a Background Site in East China. J. Meteorol. Res. 2019, 33, 905–913. [Google Scholar] [CrossRef]
  35. Li, X.X.; Chee, S.; Hao, J.M.; Abbatt, J.P.D.; Jiang, J.K.; Smith, J.N. Relative humidity effect on the formation of highly oxidized molecules and new particles during monoterpene oxidation. Atmos. Chem. Phys. 2019, 19, 1555–1570. [Google Scholar] [CrossRef]
  36. Zhang, G.Q.; Fu, H.B.; Chen, J.M. Effect of relative humidity and the presence of aerosol particles on the alpha-pinene ozonolysis. J. Environ. Sci. 2018, 71, 99–107. [Google Scholar] [CrossRef] [PubMed]
  37. Hinks, M.L.; Montoya-Aguilera, J.; Ellison, L.; Lin, P.; Laskin, A.; Laskin, J.; Shiraiwa, M.; Dabdub, D.; Nizkorodov, S.A. Effect of relative humidity on the composition of secondary organic aerosol from the oxidation of toluene. Atmos. Chem. Phys. 2018, 18, 1643–1652. [Google Scholar] [CrossRef]
  38. Lewandowski, M.; Jaoui, M.; Offenberg, J.H.; Krug, J.D.; Kleindienst, T.E. Atmospheric oxidation of isoprene and 1,3-butadiene: Influence of aerosol acidity and relative humidity on secondary organic aerosol. Atmos. Chem. Phys. 2015, 15, 3773–3783. [Google Scholar] [CrossRef]
  39. Kristensen, K.; Cui, T.; Zhang, H.; Gold, A.; Glasius, M.; Surratt, J.D. Dimers in alpha-pinene secondary organic aerosol: Effect of hydroxyl radical, ozone, relative humidity and aerosol acidity. Atmos. Chem. Phys. 2014, 14, 4201–4218. [Google Scholar] [CrossRef]
  40. Li, J.J.; Wang, G.H.; Cao, J.J.; Wang, X.M.; Zhang, R.J. Observation of biogenic secondary organic aerosols in the atmosphere of a mountain site in central China: Temperature and relative humidity effects. Atmos. Chem. Phys. 2013, 13, 11535–11549. [Google Scholar] [CrossRef]
  41. Emanuelsson, E.U.; Watne, A.K.; Lutz, A.; Ljungstrom, E.; Hallquist, M. Influence of Humidity, Temperature, and Radicals on the Formation and Thermal Properties of Secondary Organic Aerosol (SOA) from Ozonolysis of beta-Pinene. J. Phys. Chem. A 2013, 117, 10346–10358. [Google Scholar] [CrossRef]
  42. Hamed, A.; Korhonen, H.; Sihto, S.L.; Joutsensaari, J.; Jarvinen, H.; Petaja, T.; Arnold, F.; Nieminen, T.; Kulmala, M.; Smith, J.N.; et al. The role of relative humidity in continental new particle formation. J. Geophys. Res. Atmos. 2011, 116, D03202. [Google Scholar] [CrossRef]
  43. von Hessberg, C.; von Hessberg, P.; Poschl, U.; Bilde, M.; Nielsen, O.J.; Moortgat, G.K. Temperature and humidity dependence of secondary organic aerosol yield from the ozonolysis of beta-pinene. Atmos. Chem. Phys. 2009, 9, 3583–3599. [Google Scholar] [CrossRef]
  44. Pommer, L.; Fick, J.; Andersson, B.; Nilsson, C. The influence of O-3, relative humidity, NO and NO2 on the oxidation of alpha-pinene and Delta(3)-carene. J. Atmos. Chem. 2004, 48, 173–189. [Google Scholar] [CrossRef]
  45. Boy, M.; Kulmala, M. Nucleation events in the continental boundary layer: Influence of physical and meteorological parameters. Atmos. Chem. Phys. 2002, 2, 1–16. [Google Scholar] [CrossRef]
  46. Flueckiger, A.C.; Snyder, C.N.; Petrucci, G.A. Nontrivial Impact of Relative Humidity on Organic New Particle Formation from Ozonolysis of cis-3-Hexenyl Acetate. Air 2023, 1, 222–236. [Google Scholar] [CrossRef]
  47. Flueckiger, A.; Petrucci, G.A. Methodological advances to improve repeatability of SOA generation in enivronmental chambers. Aerosol Sci. Technol. 2023, 57, 925–933. [Google Scholar] [CrossRef]
  48. Chu, C.W.; Zhai, J.H.; Han, Y.M.; Ye, J.H.; Zaveri, R.A.; Martin, S.T.; Hung, H.M. New Particle Formation and Growth Dynamics for alpha-Pinene Ozonolysis in a Smog Chamber and Implications for Ambient Environments. ACS Earth Space Chem. 2022, 6, 2826–2835. [Google Scholar] [CrossRef]
  49. Tillmann, R.; Hallquist, M.; Jonsson, Å.M.; Kiendler-Scharr, A.; Saathoff, H.; Iinuma, Y.; Mentel, T.F. Influence of relative humidity and temperature on the production of pinonaldehyde and OH radicals from the ozonolysis of α-pinene. Atmos. Chem. Phys. 2010, 10, 7057–7072. [Google Scholar] [CrossRef]
  50. Tajuelo, M.; Rodriguez, D.; Rodriguez, A.; Escalona, A.; Viteri, G.; Aranda, A.; Diaz-de-Mera, Y. Secondary organic aerosol formation from the ozonolysis and oh-photooxidation of 2,5-dimethylfuran. Atmos. Environ. 2021, 245, 118041. [Google Scholar] [CrossRef]
  51. Bianchi, F.; Junninen, H.; Bigi, A.; Sinclair, V.A.; Dada, L.; Hoyle, C.R.; Zha, Q.; Yao, L.; Ahonen, L.R.; Bonasoni, P.; et al. Biogenic particles formed in the Himalaya as an important source of free tropospheric aerosols. Nat. Geosci. 2021, 14, 4–9. [Google Scholar] [CrossRef]
  52. Sulo, J.; Sarnela, N.; Kontkanen, J.; Ahonen, L.; Paasonen, P.; Laurila, T.; Jokinen, T.; Kangasluoma, J.; Junninen, H.; Sipila, M.; et al. Long-term measurement of sub-3 nm particles and their precursor gases in the boreal forest. Atmos. Chem. Phys. 2021, 21, 695–715. [Google Scholar] [CrossRef]
  53. Artaxo, P.; Hansson, H.C.; Andreae, M.O.; Back, J.; Alves, E.G.; Barbosa, H.M.J.; Bender, F.; Bourtsoukidis, E.; Carbone, S.; Chi, J.S.; et al. Tropical and Boreal Forest Atmosphere Interactions: A Review. Tellus Ser. B-Chem. Phys. Meteorol. 2022, 74, 24–163. [Google Scholar] [CrossRef]
  54. Junninen, H.; Ahonen, L.; Bianchi, F.; Quéléver, L.; Schallhart, S.; Dada, L.; Manninen, H.E.; Leino, K.; Lampilahti, J.; Mazon, S.B.; et al. Terpene emissions from boreal wetlands can initiate stronger atmospheric new particle formation than boreal forests. Commun. Earth Environ. 2022, 3, 00406–00409. [Google Scholar] [CrossRef]
  55. Kammer, J.; Flaud, P.M.; Chazeaubeny, A.; Ciuraru, R.; Le Menach, K.; Geneste, E.; Budzinski, H.; Bonnefond, J.M.; Lamaud, E.; Perraudin, E.; et al. Biogenic volatile organic compounds (BVOCs) reactivity related to new particle formation (NPF) over the Landes forest. Atmos. Res. 2020, 237, 104869. [Google Scholar] [CrossRef]
  56. Liu, Y.F.; Su, H.; Wang, S.W.; Wei, C.; Tao, W.; Pöhlker, M.L.; Pöhlker, C.; Holanda, B.A.; Krüger, O.O.; Hoffmann, T.; et al. Strong particle production and condensational growth in the uppertroposphere sustained by biogenic VOCs from the canopy of the Amazon Basin. Atmos. Chem. Phys. 2023, 23, 251–272. [Google Scholar] [CrossRef]
  57. Riipinen, I.; Pierce, J.R.; Yli-Juuti, T.; Nieminen, T.; Häkkinen, S.; Ehn, M.; Junninen, H.; Lehtipalo, K.; Petäjä, T.; Slowik, J.; et al. Organic condensation: A vital link connecting aerosol formation to cloud condensation nuclei (CCN) concentrations. Atmos. Chem. Phys. 2011, 11, 3865–3878. [Google Scholar] [CrossRef]
  58. Trostl, J.; Chuang, W.K.; Gordon, H.; Heinritzi, M.; Yan, C.; Molteni, U.; Ahlm, L.; Frege, C.; Bianchi, F.; Wagner, R.; et al. The role of low-volatility organic compounds in initial particle growth in the atmosphere. Nature 2016, 533, 527–531. [Google Scholar] [CrossRef]
  59. Dada, L.; Stolzenburg, D.; Simon, M.; Fischer, L.; Heinritzi, M.; Wang, M.; Xiao, M.; Vogel, A.L.; Ahonen, L.; Amorim, A.; et al. Role of sesquiterpenes in biogenic new particle formation. Sci. Adv. 2023, 9, eadi5297. [Google Scholar] [CrossRef] [PubMed]
  60. Quelever, L.L.J.; Kristensen, K.; Jensen, L.N.; Rosati, B.; Teiwes, R.; Daellenbach, K.R.; Perakyla, O.; Roldin, P.; Bossi, R.; Pedersen, H.B.; et al. Effect of temperature on the formation of highly oxygenated organic molecules (HOMs) from alpha-pinene ozonolysis. Atmos. Chem. Phys. 2019, 19, 7609–7625. [Google Scholar] [CrossRef]
  61. Yu, S.S.; Jia, L.; Xu, Y.F.; Zhang, H.L.; Zhang, Q.; Pan, Y.P. Wall losses of oxygenated volatile organic compounds from oxidation of toluene: Effects of chamber volume and relative humidity. J. Environ. Sci. 2022, 114, 475–484. [Google Scholar] [CrossRef]
  62. Cai, R.L.; Häkkinen, E.; Yan, C.; Jiang, J.K.; Kulmala, M.; Kangasluoma, J. The effectiveness of the coagulation sink of 3-10 nm atmospheric particles. Atmos. Chem. Phys. 2022, 22, 11529–11541. [Google Scholar] [CrossRef]
  63. Hennigan, C.J.; Bergin, M.H.; Dibb, J.E.; Weber, R.J. Enhanced secondary organic aerosol formation due to water uptake by fine particles. Geophys. Res. Lett. 2008, 35, L18801. [Google Scholar] [CrossRef]
  64. Prisle, N.L.; Engelhart, G.J.; Bilde, M.; Donahue, N.M. Humidity influence on gas-particle phase partitioning of alpha-pinene + O-3 secondary organic aerosol. Geophys. Res. Lett. 2010, 37, L01802. [Google Scholar] [CrossRef]
  65. Surdu, M.; Lamkaddam, H.; Wang, D.S.; Bell, D.M.; Xiao, M.; Lee, C.P.; Li, D.; Caudillo, L.; Marie, G.; Scholz, W.; et al. Molecular Understanding of the Enhancement in Organic Aerosol Mass at High Relative Humidity. Environ. Sci. Technol. 2023, 57, 2297–2309. [Google Scholar] [CrossRef] [PubMed]
  66. Han, Y.M.; Gong, Z.H.; Ye, J.H.; Liu, P.F.; McKinney, K.A.; Martin, S.T. Quantifying the Role of the Relative Humidity-Dependent Physical State of Organic Particulate Matter in the Uptake of Semivolatile Organic Molecules. Environ. Sci. Technol. 2019, 53, 13209–13218. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, N.X.; Jorga, S.D.; Pierce, J.R.; Donahue, N.M.; Pandis, S.N. Particle wall-loss correction methods in smog chamber experiments. Atmos. Meas. Tech. 2018, 11, 6577–6588. [Google Scholar] [CrossRef]
  68. Seinfeld, J.H.; Erdakos, G.B.; Asher, W.E.; Pankow, J.F. Modeling the formation of secondary organic aerosol (SOA). 2. The predicted effects of relative humidity on aerosol formation in the alpha-pinene-, beta-pinene-, sabinene-, Delta(3)-Carene-, and cyclohexene-ozone systems. Environ. Sci. Technol. 2001, 35, 1806–1817. [Google Scholar] [CrossRef] [PubMed]
  69. Jathar, S.H.; Mahmud, A.; Barsanti, K.C.; Asher, W.E.; Pankow, J.F.; Kleeman, M.J. Water uptake by organic aerosol and its influence on gas/particle partitioning of secondary organic aerosol in the United States. Atmos. Environ. 2016, 129, 142–154. [Google Scholar] [CrossRef]
  70. Pankow, J.F.; Marks, M.C.; Barsanti, K.C.; Mahmud, A.; Asher, W.E.; Li, J.Y.; Ying, Q.; Jathar, S.H.; Kleeman, M.J. Molecular view modeling of atmospheric organic particulate matter: Incorporating molecular structure and co-condensation of water. Atmos. Environ. 2015, 122, 400–408. [Google Scholar] [CrossRef]
  71. Qin, Y.; Ye, J.; Ohno, P.; Zhai, J.; Han, Y.; Liu, P.; Wang, J.; Zaveri, R.A.; Martin, S.T. Humidity Dependence of the Condensational Growth of α-Pinene Secondary Organic Aerosol Particles. Environ. Sci. Technol. 2021, 55, 14360–14369. [Google Scholar] [CrossRef]
  72. Pankow, J.F. Organic particulate material levels in the atmosphere: Conditions favoring sensitivity to varying relative humidity and temperature. Proc. Natl. Acad. Sci. USA 2010, 107, 6682–6686. [Google Scholar] [CrossRef] [PubMed]
  73. Gong, Z.H.; Han, Y.M.; Liu, P.F.; Ye, J.H.; Keutsch, F.N.; McKinney, K.A.; Martin, S.T. Influence of Particle Physical State on the Uptake of Medium-Sized Organic Molecules. Environ. Sci. Technol. 2018, 52, 8381–8389. [Google Scholar] [CrossRef] [PubMed]
  74. Zaveri, R.A.; Shilling, J.E.; Zelenyuk, A.; Liu, J.M.; Bell, D.M.; D’Ambro, E.L.; Gaston, C.; Thornton, J.A.; Laskin, A.; Lin, P.; et al. Growth Kinetics and Size Distribution Dynamics of Viscous Secondary Organic Aerosol. Environ. Sci. Technol. 2018, 52, 1191–1199. [Google Scholar] [CrossRef] [PubMed]
  75. Zaveri, R.A.; Shilling, J.E.; Zelenyuk, A.; Zawadowicz, M.A.; Suski, K.; China, S.; Bell, D.M.; Veghte, D.; Laskin, A. Particle-Phase Diffusion Modulates Partitioning of Semivolatile Organic Compounds to Aged Secondary Organic Aerosol. Environ. Sci. Technol. 2020, 54, 2595–2605. [Google Scholar] [CrossRef] [PubMed]
  76. Kerminen, V.M.; Lehtinen, K.E.J.; Anttila, T.; Kulmala, M. Dynamics of atmospheric nucleation mode particles: A timescale analysis. Tellus B Chem. Phys. Meteorol. 2004, 56, 135–146. [Google Scholar] [CrossRef]
  77. Kürten, A.; Williamson, C.; Almeida, J.; Kirkby, J.; Curtius, J. On the derivation of particle nucleation rates from experimental formation rates. Atmos. Chem. Phys. 2015, 15, 4063–4075. [Google Scholar] [CrossRef]
  78. Charan, S.M.; Huang, Y.L.; Seinfeld, J.H. Computational Simulation of Secondary Organic Aerosol Formation in Laboratory Chambers. Chem. Rev. 2019, 119, 11912–11944. [Google Scholar] [CrossRef]
Figure 1. A schematic of the University of Vermont Environmental Chamber (UVMEC).
Figure 1. A schematic of the University of Vermont Environmental Chamber (UVMEC).
Atmosphere 15 00480 g001
Figure 2. Nmax measured under dry (tan, solid) and humid (green, hashed) conditions for SOA generated by dark ozonolysis at different ξVOCs of (a) sabinene, (b) α-terpineol, and (c) myrtenol. Each bar represents the average Nmax (n = 2). Representative error bars are shown in panel (a) in accordance with previous work [47], assuming linearity in RSD with ξVOC.
Figure 2. Nmax measured under dry (tan, solid) and humid (green, hashed) conditions for SOA generated by dark ozonolysis at different ξVOCs of (a) sabinene, (b) α-terpineol, and (c) myrtenol. Each bar represents the average Nmax (n = 2). Representative error bars are shown in panel (a) in accordance with previous work [47], assuming linearity in RSD with ξVOC.
Atmosphere 15 00480 g002
Figure 3. NPF enhancement curve for six VOCs. ξO3 = 500 ppbv in all cases. RH = 60% except for β-pinene series (RH = 30%). α-Pinene, β-pinene, and CHA data incorporated from here [29,46]. For any given ξVOC, a ΔNmax < 0 indicates an attenuation in NPF at that ξVOC, whereas a ΔNmax = −1 indicates a complete shutdown in NPF under humid conditions at that ξVOC.
Figure 3. NPF enhancement curve for six VOCs. ξO3 = 500 ppbv in all cases. RH = 60% except for β-pinene series (RH = 30%). α-Pinene, β-pinene, and CHA data incorporated from here [29,46]. For any given ξVOC, a ΔNmax < 0 indicates an attenuation in NPF at that ξVOC, whereas a ΔNmax = −1 indicates a complete shutdown in NPF under humid conditions at that ξVOC.
Atmosphere 15 00480 g003
Figure 4. (a) The four VOCs that resulted in an enhancement of NPF under humid conditions: sabinene, α-terpineol, α-pinene, and β-pinene (from left to right). (b) The two VOCs that resulted in a shutdown of NPF under humid conditions at all ξVOCs.: myrtenol and CHA (left to right).
Figure 4. (a) The four VOCs that resulted in an enhancement of NPF under humid conditions: sabinene, α-terpineol, α-pinene, and β-pinene (from left to right). (b) The two VOCs that resulted in a shutdown of NPF under humid conditions at all ξVOCs.: myrtenol and CHA (left to right).
Atmosphere 15 00480 g004
Figure 5. Enhancement factor of α-pinene-derived SOA corresponding to varying one of three particle genesis conditions: (a) ξVOC (black, ξO3 = 500 ppbv), (b) ξO3 (red, ξVOC = 10 ppbv), and (c) RH (blue, ξO3 = 500 ppbv and ξVOC = 10 ppbv). All are scaled to a ΔNmax of 11 to show the relative impact of each condition on the enhancement of NPF.
Figure 5. Enhancement factor of α-pinene-derived SOA corresponding to varying one of three particle genesis conditions: (a) ξVOC (black, ξO3 = 500 ppbv), (b) ξO3 (red, ξVOC = 10 ppbv), and (c) RH (blue, ξO3 = 500 ppbv and ξVOC = 10 ppbv). All are scaled to a ΔNmax of 11 to show the relative impact of each condition on the enhancement of NPF.
Atmosphere 15 00480 g005
Figure 6. Amplification factor of α-pinene-derived SOA for three chamber parameters: (a) ξVOC (black), (b) ξO3 (red), and (c) RH (blue). All plots are scaled to a maximum ΔE of 11 to show the relative impact of each condition on the enhancement of NPF.
Figure 6. Amplification factor of α-pinene-derived SOA for three chamber parameters: (a) ξVOC (black), (b) ξO3 (red), and (c) RH (blue). All plots are scaled to a maximum ΔE of 11 to show the relative impact of each condition on the enhancement of NPF.
Atmosphere 15 00480 g006
Figure 7. The normalized apparent rate of particle formation (Jap) of α-pinene-derived SOA, corresponding to three particle genesis conditions: (a) ξVOC (black), (b) ξO3 (red), and (c) RH (blue). All are scaled to a ΔE of 11 to show the relative impact of each condition on the enhancement of NPF.
Figure 7. The normalized apparent rate of particle formation (Jap) of α-pinene-derived SOA, corresponding to three particle genesis conditions: (a) ξVOC (black), (b) ξO3 (red), and (c) RH (blue). All are scaled to a ΔE of 11 to show the relative impact of each condition on the enhancement of NPF.
Atmosphere 15 00480 g007
Figure 8. Log–log plot of amplification factor (ΔE) vs. normalized apparent rate of particle formation (Jap) for ξVOC (black squares) and RH (blue triangles) of α-pinene-derived SOA. The slope of the RH series is approximately 16x greater than that of the ξVOC series.
Figure 8. Log–log plot of amplification factor (ΔE) vs. normalized apparent rate of particle formation (Jap) for ξVOC (black squares) and RH (blue triangles) of α-pinene-derived SOA. The slope of the RH series is approximately 16x greater than that of the ξVOC series.
Atmosphere 15 00480 g008
Table 1. ΔNmax for each VOC at varying ξVOCs. Dashes indicate that ΔNmax could not be calculated at that ξVOC because no NPF was observed under dry conditions. ΔNmax experiments were not conducted for sabinene and myrtenol at 20 ppbv because it had already stabilized at lower ξVOC. Error in ξVOC estimated by propagating errors in air flow rate to chamber and syringe error. Errors in ΔNmax were estimate from the RSDs interpolated for the respective Nmax,0RH based on a previous study assessing the repeatability of SOA experiments in the UVMEC [47].
Table 1. ΔNmax for each VOC at varying ξVOCs. Dashes indicate that ΔNmax could not be calculated at that ξVOC because no NPF was observed under dry conditions. ΔNmax experiments were not conducted for sabinene and myrtenol at 20 ppbv because it had already stabilized at lower ξVOC. Error in ξVOC estimated by propagating errors in air flow rate to chamber and syringe error. Errors in ΔNmax were estimate from the RSDs interpolated for the respective Nmax,0RH based on a previous study assessing the repeatability of SOA experiments in the UVMEC [47].
ξVOC (ppbv)Sabineneα-TerpineolMyrtenol
0.20 (±0.02)2.88 (±0.25)----
0.35 (±0.03)0.47 (±0.04)----
0.50 (±0.03)−0.02 (±0.001)--−0.77 (±0.06)
1.00 (±0.03)−0.02 (±0.001)--−0.64 (±0.02)
1.5 (±0.03)--9.77 (±0.86)--
4.0 (±0.1)−0.64 (±0.02)2.35 (±0.07)−0.66 (±0.02)
10.0 (±0.1)−0.39 (±0.01)0.04 (±0.001)−0.85 (±0.03)
20.0 (±0.3)--−0.09 (±0.003)--
Table 2. The VOCs addressed and their corresponding solubility. * Solubilities as reported in https://pubchem.ncbi.nlm.nih.gov/ (accessed on 23 February 2024).
Table 2. The VOCs addressed and their corresponding solubility. * Solubilities as reported in https://pubchem.ncbi.nlm.nih.gov/ (accessed on 23 February 2024).
VOCSolubility in H2O (mg/L) *
Sabinene2.5
α-Terpineol7100
Myrtenol427
α-PineneInsoluble
β-PineneInsoluble
CHA900
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Flueckiger, A.C.; Petrucci, G.A. Effect of Relative Humidity on the Rate of New Particle Formation for Different VOCs. Atmosphere 2024, 15, 480. https://doi.org/10.3390/atmos15040480

AMA Style

Flueckiger AC, Petrucci GA. Effect of Relative Humidity on the Rate of New Particle Formation for Different VOCs. Atmosphere. 2024; 15(4):480. https://doi.org/10.3390/atmos15040480

Chicago/Turabian Style

Flueckiger, Austin C., and Giuseppe A. Petrucci. 2024. "Effect of Relative Humidity on the Rate of New Particle Formation for Different VOCs" Atmosphere 15, no. 4: 480. https://doi.org/10.3390/atmos15040480

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop