Introduction

Dark-coloured aerosols such as combustion-induced carbonaceous particles and wind-blown mineral dust absorb solar radiation and perturb the climate system by heating the atmosphere and reducing the snow albedo1,2. Shortwave atmospheric heating by such aerosols can alter the cloud fraction and hydrological cycle on both regional and global scales3,4,5,6. The global mean increase in precipitation per degree of increase in global-mean surface temperature, attributable to human-induced global warming, strongly depends on the emission of black carbon (BC), a proxy for anthropogenic shortwave absorbers6. To evaluate the effects of dark-coloured aerosols on climate using numerical models, observational datasets are needed to constrain the sources, atmospheric abundance, and detailed microphysical properties (for example, size distribution and morphology) of individual light-absorbing aerosols1,7. The long-term data on absorption aerosol optical depth measured by ground-based remote sensing8 have been the main observational constraints for numerical models used to evaluate the effects of absorbing aerosols on climate change1,7. Inverting the remote sensing data to the column abundance of absorbing aerosols requires a priori assumption on the optical properties of each type of aerosols that contributes to absorption aerosol optical depth9. Therefore, in situ observations of all the important contributors to atmospheric shortwave absorption form the basis for the quantitative investigation of the complex effects of absorbing aerosols on climate.

Until now, two types of carbonaceous aerosols, BC and brown carbon (BrC), and wind-blown mineral dust (DU) have generally been considered as the only three aerosol species that contribute significantly to shortwave absorption in the atmosphere and snowpack1,2,8,10,11. BC and BrC are mostly co-emitted during the burning of fossil fuels, biofuels and biomass7. Light absorption by DU is mostly due to iron oxide nanoparticles attached on the surfaces of host mineral materials12. DU absorption dominates the atmospheric shortwave absorption over the desert and dust outflow regions13. A number of attempts have recently been made to quantify BrC absorption separately from BC absorption11,14,15,16. These studies suggest that BC absorption almost invariably dominates BrC absorption, except at ultraviolet wavelengths. In Beijing, an indirect estimate using ground-based remote-sensing technique14 indicated that BrC absorption is 10% of BC absorption at the mid-visible wavelength in the winter and spring seasons. Another estimate of BrC absorption in Beijing using in situ optical and chemical measurements in March provided a similar result11. Using a global chemical transport model, Feng et al. estimated that BrC accounts for 7–19% of aerosol absorption (global average)17.

Recent observational studies using electron microscopy methods revealed that anthropogenic iron oxide particles in the form of aggregated FeOx nanoparticles are ubiquitous in urban atmospheres18,19,20 and roadside environments21. They are emitted from, for example, the blast furnaces of iron manufacturing facilities22 and from the engines and brakes of motor vehicles23,24. The major iron oxide phases of aggregated FeOx nanoparticles are magnetite (Fe3O4)18,20,21,23,24, maghemite (γ-Fe2O3)21,24 and hematite (α-Fe2O3)18,21,24. However, their abundances in the troposphere and radiative effects have not yet been evaluated.

In this study, in situ aircraft measurements using a modified single-particle soot photometer (SP2)25 and electron microscopy are performed to show that anthropogenic FeOx particles, particularly aggregated magnetite nanoparticles, are ubiquitous in the continental outflows from East Asia. We then evaluate their contribution to atmospheric shortwave absorption on the basis of the observed size-resolved mass concentrations and particle morphologies. Our results indicate that the absorption by anthropogenic FeOx is at least 4–7% of the BC absorption over East Asia.

Results

Observation

We used observational data from the Aerosol Radiative Forcing in East Asia (A-FORCE) 2013W aircraft campaign26 over Yellow Sea and East China Sea in February and March, 2013. Our modified SP2 can measure individual BC and FeOx particles in the mass-equivalent diameter (Dm) domains of 70 nm≤Dm≤850 nm and 170 nm≤Dm≤2,100 nm, respectively25. The aerosol-sampling system, which consisted of a forward-facing shrouded solid diffuser inlet, transport tubes and aerosol measuring instruments, was designed to observe submicron-sized particles. The theoretical transmission efficiency curves Tr(Dm) of FeOx particles began to drop at Dm=600 nm and decreased to 0.5 at Dm=2,100 nm (Supplementary Fig. 1). These theoretical Tr(Dm) curves suggest that our measurement system underestimates the ambient FeOx concentrations at Dm>600 nm.

Characterization of anthropogenic FeOx

In this section, we characterize the fundamental properties of airborne FeOx particles in the East Asian continental outflows. For this purpose, we focus on the air parcel passed through the planetary boundary layer over eastern China and was directly transported to the sampling point on the flight track below 2 km altitude without experiencing wet removal of aerosols. The aircraft observation data of this air parcel is called dry PBL air26. The detailed criteria for selecting the dry PBL air according to the backward trajectory analysis is described in Kondo et al.26.

First, we classified individual Fe-bearing particles in dry PBL air depending on the composition and morphology, based on the electron-microscopy analyses of 1,460 particles collected by an onboard aerosol-impactor sampler. A transmission electron microscope (TEM) and a scanning transmission electron microscope equipped with energy-dispersive X-ray spectrometry (STEM–EDS)19 were used for these analyses. Table 1 lists the number of Fe-bearing particles for each of the three morphological types measured by the STEM–EDS analyses. The most abundant type of Fe-bearing particle was aggregated FeOx nanoparticles; the TEM images and associated elemental mappings of these nanoparticles are shown in Fig. 1; Supplementary Fig. 2. The diameters of the FeOx monomers comprising each aggregate were highly variable, ranging from several nm to 100 nm. These composition and morphology of FeOx particles are similar to those found in Tokyo19 or Mexico City27. Our electron energy loss spectroscopy (EELS) analyses with TEM showed that the major component of the aggregated FeOx nanoparticles in dry PBL air is magnetite (Supplementary Fig. 3). The second most abundant type of Fe-bearing particles in dry PBL air was fly ash, which is a complex internal mixture of combustion-induced refractory materials commonly includes Si as a major component. The Fe-bearing fly ash particles in the dry PBL air typically contained only several per cent of Fe by mass. A third type of Fe-bearing particle, Fe-bearing mineral dust, was rare in the dry PBL air; the fraction of mineral dust particles among all the Fe-bearing particles was only 0.035 (Table 1). Although the mineral dust was observed to be minor in the dry PBL air, it would be the dominant type of Fe-bearing particles over the East Asia during an Aeolian dust (that is, Kosa) outflow event.

Table 1 Number of particles determined by STEM–EDS analysis for samples collected in dry PBL air.
Figure 1: Transmission electron microscopy images and element distributions of aggregated iron oxide nanoparticles found in dry PBL air.
figure 1

The sample was collected using an aerosol-impactor sampler onboard the aircraft from 14:35 to 14:47 on 4 March 2013 (local time) during the A-FORCE 2013W campaign (that is, Sample number 1 in Table 1). The particles magnified in a and b were collected on the same substrate with an inter-particle distance of 4 μm, as shown in c. The elemental distributions of Fe and O for particles a and b are shown in d and e, respectively, and indicate that these particles contain iron oxide.

Next, we present results provided by the light-scattering signals of individual FeOx particles measured by the SP2. Figure 2 shows a scatterplot of the scattering cross-section at the onset of the incandescence (Cs-oi)28 and the mass-equivalent diameter Dm for both laboratory and ambient FeOx-containing particles. Figure 2 also includes experimental results for pure magnetite particles and mineral dust particles (Icelandic dust and Taklamakan dust)25. Compared with pure magnetite, the mineral dust samples exhibited large Cs-oi values beyond the detectable limit of 4 × 10−14 m2, because the host minerals internally mixed with the incandescing FeOx also contribute to the Cs-oi value. On the basis of these results, we introduced a criterion for classifying the mixing state of detected FeOx-containing particles in the Dm domain of 170 nm≤Dm≤270 nm. Particles with Cs-oi>2 × 10−14 m2 were classified as dust-like. Figure 2 shows that a majority of the FeOx-containing particles in the dry PBL air were similar to the pure magnetite particles in terms of their Cs-oi value, whereas a minority had substantially larger Cs-oi values. In the dry PBL air, the fraction of dust-like FeOx particles among the detected FeOx-containing particles was 0.1, as discussed later.

Figure 2: Single-particle soot photometer data indicating the mixing state of individual iron oxide-containing particles.
figure 2

The figure shows scatterplots of scattering cross-section at the onset of incandescence (Cs-oi) and mass-equivalent diameter Dm for laboratory samples (magnetite, Icelandic dust and Taklamakan dust) and FeOx-containing particles in dry PBL air. For dry PBL data, the solid red line and two dashed red lines represent median and (10 and 90) percentile Cs-oi values as functions of Dm, respectively. The Dm domain used for classifying the mixing state of FeOx-containing particles is schematically shown. In this study, ambient FeOx-containing particles with 170 nm≤Dm≤270 nm are classified as dust-like FeOx if Cs-oi > 2 × 10−14 m2. The detectable upper limit of Cs-oi is 4 × 10−14 m2.

In addition to the direct microscopic observations using the TEM–EELS, the optical signals measured using the SP2 also provided indirect information on the material properties of the airborne FeOx. The timing of the onset of incandescence (toi) of FeOx-containing particles has been experimentally shown to be a qualitative indicator of colour darkness (that is, light-absorbing efficiency) in FeOx materials25. The distribution of the toi values of the FeOx-containing particles in dry PBL air was similar to that for black-coloured magnetite but dissimilar to that of red-coloured hematite (see the Methods section and Supplementary Fig. 4). This implies that the major component of incandescing FeOx particles in dry PBL air is magnetite, which is consistent with the TEM–EELS results (Supplementary Fig. 3).

We found that the FeOx number concentration in the dry PBL air was highly correlated with the CO mixing ratio (r2=0.87) and BC number concentration (r2=0.69; Fig. 3a,b, respectively). These results suggest that the spatial distribution of the FeOx emission flux over the East Asian continent is similar to those of CO and BC. On the other hand, the FeOx/BC number concentration ratio in the dry PBL air-correlated neither with the relative abundance of the supermicron-sized aerosols (Fig. 3c) nor the number fraction of the dust-like FeOx-containing particles (Fig. 3d). The scatterplots imply that the majority of the detected FeOx particles in the dry PBL air are not associated with DU, which is consistent with the results of the direct STEM–EDS analyses of the aerosol-impactor samples (Table 1).

Figure 3: Scatterplots between the 1-min data of various observed parameters in dry PBL air.
figure 3

Number of data points (n) and the square of the correlation coefficient (r2) are shown in each panel. (a,b) The correlations between the FeOx concentrations and the concentrations of two representative pollutants in urban air: BC and CO. Particle concentrations were calculated at standard temperature and pressure (STP). The red line represents the linear regression line. (c,d) Scatterplots between the FeOx/BC concentration ratio and two available indicators of relative abundance of mineral dust particles: (c) number concentration ratio of supermicron-sized aerosols to submicron-sized aerosols and (d) number fraction of dust-like FeOx particles. In c, Np denotes the number concentration of aerosols within a particular range of light-scattering equivalent diameter Dp, which was observed using a cloud and aerosol spectrometer probe.

The two most important conclusions drawn from these observations are as follows. First, the FeOx particles detected using the modified SP2, with the exception of the dust-like FeOx, are primarily aggregated magnetite nanoparticles of anthropogenic origin. Second, the anthropogenic magnetite is a major type of Fe-bearing aerosol in East Asian continental outflow.

Altitude profiles

Figure 4 shows the size-resolved number and mass concentrations of FeOx at each altitude along with those in the dry PBL air. Within the observed size domain of 170 nm≤Dm≤2,100 nm, the size-resolved number concentration of FeOx was approximated using a power function with an offset

Figure 4: Size-resolved number and mass concentrations of iron oxide particles.
figure 4

Results in dry PBL air and at different altitudes are shown in a and bf, respectively. Filled circle and error bar represent the sample mean and the sample s.d. (±1σ) of the 1-min data, respectively. The number of 1-min data points (n) is shown in each panel. Dashed lines show the results of least-square fitting of equation (1).

where y0, a and p are numerical parameters listed in Table 2. The size-resolved mass concentration of FeOx was approximated as

Table 2 List of the parameters for the size-distribution functions (equations (1) and (3)).

where is the assumed bulk density of FeOx (5.17 g cm−3). In contrast, the size-resolved mass concentration of BC within the observed size domain of 70 nm≤Dm≤850 nm (see Figs 6 and 14 of Kondo et al.26) was approximated with a two-modal lognormal function as

where Ai, Di and σgi (i=0,1) are numerical parameters listed in Table 2. Figure 4 shows that the mode Dm values of the number size distribution were smaller than 170 nm, whereas the mode Dm values of the mass size distribution may have been larger than 2,100 nm, that is, the detectable Dm domain of 170 nm≤Dm≤2,100 nm was insufficient to reveal the entire shape of the size-distribution function of ambient FeOx. As mentioned earlier, our aerosol-sampling system likely underestimated the ambient FeOx concentrations for Dm>600 nm. In contrast, our BC measurements in the domain of 70 nm≤Dm≤850 nm were likely sufficient to determine the total BC mass concentrations, as expected from the Figure 6 and 14 of Kondo et al.26 and the parameters listed in Table 2.

The BC size distributions tended to shift towards smaller sizes at higher altitudes26 (see parameter D0 in Table 2). This trend has been explained by the preferential wet removal of larger BC particles during vertical transport26,29. In contrast, the FeOx size distributions shifted towards larger sizes with increasing altitude (Fig. 4). The decrease of the parameter p with altitude reflects this trend (Table 2). As shown later, the number fraction of dust-like FeOx was observed to increase with altitude. We hypothesize that the mineral dust particles lifted from the deserts in central China and transported by the westerlies cause the larger shift of the FeOx size distributions over the Yellow Sea and East China Sea.

Figure 5a,b show the vertical profiles of the FeOx concentrations and the FeOx/BC concentration ratios, respectively. The mean FeOx mass concentration was 100–400 ng m−3 at standard temperature and pressure (STP) in the boundary layer (altitude<2 km), and monotonically decreased with altitude to 30–90 ng m−3 STP in the free troposphere (altitude>3 km). The mean FeOx/BC number concentration ratio was 0.004 in the boundary layer and decreased to 0.002 in the free troposphere. The mean FeOx/BC mass concentration ratio was 0.3–0.5 in all altitudes below 6 km, and increased to 0.8 at higher altitudes (6–8 km). The observed FeOx/BC number (mass) ratio would be substantially larger if we were able to detect smaller (larger) FeOx particles with Dm<170 nm (Dm>2,100 nm). The number fraction of dust-like FeOx showed a monotonic increase with altitude from 0.07 at 0–1 km to 0.5 at 6–8 km (Fig. 5c). In an aerosol-impactor sample collected during a time period of the highest FeOx number concentration observed at 6–8 km altitude (0.4 cm−3 STP), we actually found mineral dust particles including Fe (Supplementary Fig. 5). However, in the free tropospheric samples, the number of Fe-bearing particles collected on each TEM grid was too low to evaluate the relative abundance of each of the three morphological types listed in Table 1.

Figure 5: Altitude profiles of various observed parameters.
figure 5

In each panel, the filled circles represent the mean values at each altitude, whereas the filled triangle represents the mean value in dry PBL air. (a) FeOx mass and number concentrations. The error bars in a represent the sample standard deviation (±1σ) of 1-min data. (b) The mean FeOx/BC concentration ratios for mass and number. (c) The number fraction of dust-like FeOx for FeOx particle sizes in the Dm domain of 170 nm≤Dm≤270 nm. The error bars in c represent uncertainty estimated by assuming that the number of particles detected in each altitude range follows a Poisson distribution. (d) The log(Cs-be/Cs-oi) value of BC-containing particles for BC particle sizes in the Dm domain of 205 nm≤Dm≤220 nm. The error bars in d represent the ±1σ ranges of single-particle data.

Mass-absorption cross-sections

An original discrete-dipole approximation code was used to calculate the mass-absorption cross-sections of the BC- and FeOx-containing particles by assuming the particle shape, mixing state and refractive index of each material. Before discussing the results, we briefly explain the nontrivial assumptions used. For both BC and FeOx, the particle shape was assumed to be a fractal-like aggregate of spherical monomers. The mixing state of the aggregate with other non-absorbing materials was assumed to be either bare or coated: the bare state denotes a pure aggregate, whereas the coated state denotes an aggregate coated by a large amount of non-absorbing material. In the coated state, the volume of the coating material was prescribed such that the theoretical log(Cs-be/Cs-oi) value of the model BC-containing particles was greater than the measured log(Cs-be/Cs-oi) values of the real BC-containing particles, where the Cs-be denotes the scattering cross-section of a BC-containing particle in the SP2 laser beam before the onset of evaporation28 (see the Methods section). The mean+1σ value of the observed log(Cs-be/Cs-oi) value for BC-containing particles with Dm 200 nm was 0.9 at all altitudes below 8 km (Fig. 5d). On the basis of this observation, the coating/aggregate volume ratio in the coated state was chosen to be approximately 3–4 so that the theoretical value of log(Cs-be/Cs-oi) for the coated BC was 1.0 at Dm 200 nm (Supplementary Fig. 6).

The mass-absorption cross-sections (σa) calculated for the BC- and FeOx-containing particles as functions of wavelength (λ) and mass-equivalent diameter Dm are shown in Fig. 6. In general, the σa value at a particular (λ, Dm) is approximately three times larger for BC than for FeOx, primarily because the assumed bulk density of FeOx (5.17 g cm−3) was 2.87 times greater than that of BC (1.8 g cm−3). Compared with the bare state, σa was enhanced by a factor of approximately two in the coated state because of the so-called lensing effects30. The σa(λ, Dm) of FeOx did not change appreciably with λ in the Dm domain largely contributing the total ambient FeOx mass concentration. Thus, we expect that the atmospheric absorption coefficient (babs) attributable to anthropogenic FeOx depends little on the wavelength. In contrast, the babs attributable to carbonaceous aerosols (BC+BrC) is known to decrease sharply with wavelength14,15,31. This large difference in the wavelength dependence of babs between anthropogenic FeOx and carbonaceous aerosols will be useful to classify them in remote sensing observations.

Figure 6: Mass-absorption cross-sections of black carbon- and iron oxidex-containing particles as a function of wavelength and mass-equivalent diameter.
figure 6

The mass-absorption cross-section σa(λ, Dm) (m2 g−1) are computed for four distinct particle types: (a) bare BC, (b) bare FeOx, (c) coated BC, and (d) coated FeOx.

Shortwave atmospheric heating

Here we quantify the instantaneous shortwave atmospheric heating attributable to FeOx and BC using their observed size-resolved mass concentrations and theoretical mass-absorption cross-sections. The contribution to shortwave atmospheric heating from particles of a particular size is quantified by the size-resolved absorbing power dPabs/dlogDm (W m−3), which is defined as

where Fac(λ) is the spectral actinic flux (W m−2 nm−1), σa(λ, Dm) is the mass-absorption cross-section (m2 g−1) and dM/dlogDm (g m−3) is given by equations (2) or (3). The value of Fac(λ) for each altitude was computed using a radiative transfer model assuming a clear sky and a daytime-mean solar zenith angle (see the Methods section). Figure 7 shows the size-resolved absorbing powers of BC and FeOx in (a) dry PBL air and (b) at the highest altitude (6–8 km). The bell-shaped dPabs/dlogDm distributions for BC suggest that the BC-containing particles within the detectable Dm domain (70 nm≤Dm≤850 nm) predominantly contribute to the total BC absorption. On the other hand, the broad dPabs/dlogDm distributions of FeOx suggest that the unobserved FeOx-containing particles outside the detectable Dm domain (170 nm≤Dm≤2,100 nm) also contribute significantly to the total FeOx absorption. Figure 7 also shows the theoretical transmission efficiency curves (Tr(Dm)) of our aerosol-sampling system for BC and FeOx particles. The steeply decreasing Tr(Dm) curve for FeOx suggests that we underestimated the dPabs/dlogDm value for FeOx at Dm>600 nm.

Figure 7: Size-resolved shortwave absorbing powers dPabs/dlogDm of black carbon and iron oxide.
figure 7

The dPabs/dlogDm values (thin lines) were calculated for two distinct atmospheric conditions: (a) dry PBL air (1 km altitude) and (b) the highest altitude (6–8 km). Solid and dashed lines represent the mean dPabs/dlogDm values for bare and coated particles, respectively. Error bars represent the variability in the dPabs/dlogDm values estimated from the ±1σ value of the 1-min data of dM/dlogDm. In each panel, the transmission efficiency curves Tr(Dm) of the aerosol-sampling system calculated for BC and FeOx are also shown (thick lines).

The total shortwave absorbing power Pabs (W m−3) was computed by integrating equation (4) over Dm:

where Dmin and Dmax denote the lower and upper limits of the detectable Dm domain, respectively. Figure 8a shows the altitude profiles of Pabs(BC) and Pabs(FeOx). For both BC and FeOx, Pabs values decreased monotonically with altitude, following their mass concentration profiles. For the dry PBL air, the mean Pabs(BC) and Pabs(FeOx) values in the bare (coated) state were 4.0 (10.0) and 0.24 (0.44) mW m−3, respectively. These Pabs values are equivalent to atmospheric heating rates of 0.29 (0.72) and 0.017 (0.032) K per day, respectively. Figure 8b shows the altitude profiles of the mean Pabs(FeOx)/Pabs(BC) ratio. In the boundary layer wherein anthropogenic FeOx dominated the detected FeOx particles (Fig. 5c), the mean Pabs(FeOx)/Pabs(BC) ratio was estimated to be 0.04–0.07. The altitude dependence of Pabs(FeOx)/Pabs(BC) was similar to that of the FeOx/BC mass concentration ratio (Fig. 5b). In the highest altitude (6–8 km), wherein the dust-like FeOx particles contributed 50% of the detected FeOx particles (Fig. 5c), the Pabs attributable to anthropogenic FeOx may be 50% of the calculated Pabs value.

Figure 8: Altitude profiles of the shortwave absorbing powers Pabs attributable to black carbon and iron oxide.
figure 8

(a) The Pabs(BC) and Pabs(FeOx), whereas b shows mean Pabs(FeOx)/Pabs(BC) ratio. Filled and open circles (triangles) represent the mean values in each altitude range (dry PBL air) for the bare and coated particles, respectively. In a, error bars represent the variability in Pabs estimated from the ±1σ value of the 1-min mass concentration data.

It should be noted that the Pabs(FeOx)/Pabs(BC) ratios reported herein are lower than the real values because our measurements underestimated the ambient FeOx mass concentration in the Dm domain of 600 nm≤Dm≤2,100 nm and totally ignored the FeOx particles outside the detectable Dm domain. Considering these facts, the real Pabs(FeOx)/Pabs(BC) ratios in the boundary layer would be as large as 0.1.

Discussion

On the basis of the previous observations of BrC near Beijing in March 2013 (refs 11, 14), we roughly estimate the typical Pabs(BrC)/Pabs(BC) value in the boundary layer to be 0.1. Thus, we expect that Pabs(FeOx) is as large as Pabs(BrC) in the East Asian continental outflows. In future studies, it is highly desirable to use an aerosol-sampling system with a higher transmission efficiency of large FeOx particles. In addition, some technical improvements in the SP2 are necessary to perform FeOx measurements beyond the current detectable Dm domain (170 nm≤Dm≤2,100 nm). Despite the room for such improvements, an essential conclusion can be drawn from our results, namely, in addition to BC and BrC, airborne FeOx in the form of aggregated magnetite nanoparticles should also be recognized as a significant anthropogenic contributor to shortwave atmospheric heating.

In addition to clear-sky shortwave absorption, we briefly discuss some other potential climate effects of anthropogenic FeOx particles. Low stratiform clouds are of frequent climatological occurrences over the mid- to high-latitude ocean and southeast China32. Under modest maximum supersaturation in such clouds, particle’s mass fraction activated to cloud droplets will be substantially larger for FeOx than BC, because the critical supersaturations of FeOx-containing aerosols will be lower than those of BC due to the larger Dm. Under these situations, Pabs(FeOx)/Pabs(BC) is enhanced because the lensing effect in droplets33 is larger for FeOx than for BC. We expect that the number of co-emitted cloud condensation nuclei and their precursor gases in anthropogenic FeOx-rich sources, which remain uncertain, are substantially different from those in anthropogenic BC- and BrC-rich sources such as residential coal, industrial coal and biomass fuels7. Thus, without comprehensive investigations, it is not clear whether the net positive climate forcing of anthropogenic FeOx-rich sources is negligible or comparable with those of BC- and BrC-rich sources. Finally, it should be mentioned that the anthropogenic FeOx particles may also play a role in the biogeochemical cycles34.

Methods

Modified single-particle soot photometer

A modified single-particle soot photometer (SP2), which detects light-absorbing refractory aerosols on the basis of intra-cavity laser-induced incandescence, was used to measure BC and iron oxide FeOx particles25. The BC and FeOx are discriminated from each other on the basis of the colour ratio, which is an indicator of the boiling temperature25, and the peak amplitude of the blue-band incandescence signal, which is an indicator of an incandescing particle’s size25. Supplementary Figure 7 shows a scatterplot of the peak amplitudes and colour ratios of all incandescing particles detected during the A-FORCE 2013 W campaign. The boundary lines for discriminating FeOx from BC were determined on the basis of experimental results25 and are also shown in Supplementary Fig. 7.

The masses (m) of the individual FeOx and BC particles were determined from the peak heights of the incandescence signals using the experimentally determined mass-to-peak height relationships25. The mass-equivalent diameters (Dm) for BC and FeOx were calculated from the observed masses m assuming the void-free densities of 1.8 g cm−3 and 5.17 g cm−3 (density of magnetite), respectively. In this study, the detectable size domains of the BC and FeOx particles were 70 nm≤Dm≤850 nm and 170 nm≤Dm≤2100, nm, respectively.

The detailed morphological properties of the individual particles containing a particular mass of each incandescing material were evaluated on the basis of time-resolved scattering cross-sections Cs (integrated over the solid angle of light collection) in a laser beam derived from the scattering signal35,36,37. We evaluated whether the incandescing material (BC or FeOx) was attached to the surface of another particle on the basis of the magnitude of Cs at the onset of the incandescence signal (Cs-oi)28. In our observations, the fraction of the attached-type28 BC-containing particles with Dm=200 nm was <3% at any altitudes below 8 km. Thus, we assumed the Cs-oi of BC-containing particles to be equivalent to the Cs of BC core. For BC-containing particles, the ratio of Cs before the onset of particle evaporation (Cs-be) to Cs-oi is used as an indicator of the amount of non-refractory material (for example, sulfate) coating the BC28.

The timing of the onset of the incandescence signal (toi) is an indicator of the heating rate of an absorbing particle in the laser beam of the SP2; toi tends to be lower (earlier) for more efficient light absorbers. Our experiments25 showed that the toi of magnetite (Fe3O4) particles was markedly earlier than that of hematite (α-Fe2O3), reflecting the greater absorption efficiency of magnetite. It should be noted that comparisons of toi values derived from the laboratory experiments and field data are meaningful only under similar SP2 conditions (that is, the same laser power and width of the Gaussian beam) because the toi of a particular composition depends on these parameters. Supplementary Fig. 4 shows the toi distributions for the ambient BC and FeOx particles in dry PBL air. The SP2 laser power during the A-FORCE 2013W campaign was similar to that in our laboratory experiments25, as expected from the comparison of toitcen (0.2 μs) for BC between ambient data (−55<toitcen<−30) and our laboratory results (−50<toitcen<−35; refer to Fig. 2c of Yoshida et al.25). The toi distributions of the FeOx particles for two different mass ranges (10 fg≤m≤88 fg and m > 530 fg) were similar, indicating that the primary FeOx material does not change appreciably with FeOx particle mass (Supplementary Fig. 4). Our experiment illustrated that the incandescing probability of pure hematite particles was zero for m<100 fg because the absorption efficiency of hematite particles in this size range is insufficient to heat the particles to the incandescing temperature25. Thus, the incandescing FeOx particles in dry PBL air had greater absorption efficiency than hematite. The toitcen (0.2 μs) distributions of FeOx with 10 fg≤m≤88 fg and m>530 fg largely overlapped with the toitcen distributions of pure magnetite with m <100 fg (−40<toitcen<−20)25 and m>530 fg (−40<toitcen<−30)25, respectively. On the basis of these results and the EELS spectra shown in Supplementary Fig. 3, the major constituent of the detected FeOx particles in dry PBL air is likely magnetite (Fe3O4).

Electron microscopy analysis

An aerosol impactor-sampler19 onboard the aircraft was used to collect aerosol samples on the Cu TEM grids with collodion substrate at 12-min intervals during each flight. A 120-kV transmission electron microscope (JEM-1400, JEOL) equipped with an energy-dispersive X-ray spectrometer (Oxford Instruments) was used for the STEM–EDS analysis. A 200-kV transmission electron microscope (ARM 200, JEOL) was used for the EELS analysis.

Transmission efficiency of aerosols

Special care is required to measure the number concentrations of supermicron-sized aerosols in ambient air as the transmission efficiency (Tr) through the entire tubing apparatus connecting the aerosol inlet to the particle detection volume unavoidably decreased because of the inertial and gravitational depositions. Since a substantially large fraction of the total FeOx mass was expected to be in the supermicron size range (Fig. 4), our conclusion is strongly dependent on the degree of sampling loss of large FeOx particles. According to the theoretical formulae compiled by Pramod et al.38, we estimated the Tr(Dm) curves of the BC and FeOx particles through our aerosol-sampling system. In our A-FORCE 2013W aircraft campaign, the sample air was aspirated through a forward-facing shrouded inlet (DMT Inc., Boulder, CO, USA) installed on the top of the aircraft fuselage. This aerosol inlet is a replica of ‘University of Hawaii shrouded solid diffuser inlet’ described and evaluated by McNaughton et al.39. An isokinetic aspiration was performed to maintain the aspiration efficiency at 1.0, independent of the particle size. The geometric specifications and flow rates of the tubing apparatuses comprising the sampling system are listed in Supplementary Table 1. The void-free densities of BC and FeOx were assumed to be 1.8 and 5.17 g cm−3, respectively. The dynamic shape factors of the BC and FeOx particles were assumed to be 1.5. Our experiment using pure magnetite particles confirmed the reasonable agreement between the measured and theoretical Tr(Dm) curves for a ¼-inch tube (length=0–3 m) assuming the same density and dynamic shape factor.

Computing the fractal-like aggregates of spheres

On the basis of the TEM observations of ambient BC- and FeOx-particles, we assumed the shapes of the model BC- and FeOx-particles for the electromagnetic scattering calculations to be fractal-like aggregates of spheres. The aggregate geometry was computed using an original tunable cluster–cluster aggregation (CCA) code called aggregate_gen, which is an efficient C++ implementation of the hierarchical CCA algorithm40. We assumed the fractal prefactor and fractal dimension to be kf=1.0 and Df=2.8, respectively. The monomer diameters of BC and FeOx were assumed to be 40 and 80 nm, respectively. The number of monomers Npp in the aggregates ranged from 8 to 16384. This range of Npp covers the observed Dm domain for both BC and FeOx.

Electromagnetic scattering calculations

The mass-absorption cross-sections for BC and FeOx as functions of both wavelength λ and mass-equivalent diameter Dm were computed using an original electromagnetic scattering solver called block-DDA, which utilizes the block Krylov subspace methods41,42,43 to efficiently solve the discrete-dipole approximation (DDA)44,45 for multiple incident waves, or equivalently, for multiple target orientations. For every (λ, Dm) condition, we used the mean value of the results from four different randomly chosen target orientations. When applying block-DDA to the scattering problems for fractal-like aggregates of BC or FeOx monomers, the monomer dipole formulation was assumed to avoid the discretization shape error in each monomer46. The coupled electric and magnetic dipoles formulation46,47 was used to mitigate the multipole-truncation error associated with the monomer-dipole assumption46. To apply efficient fast Fourier transform-based algorithms in DDA48, we adjusted the center position of each monomer dipole to the nearest site of the computational cubic lattice (CL)46. In this study, the lattice spacing of the CL was set to one-half of the monomer diameter. The geometry of the coating material surrounding a fractal-like aggregate of spheres was computed by the following procedure. First, the CL sites in the proximity of each monomer (at a distance of less than T monomer radii) were selected as candidates for coating volume elements, where the coating thickness was controlled by the parameter T. In this study, T was selected to be 3. Next, the CL sites overlapping with monomer volume, if present, were removed from the candidates. Finally, the coating volume was assigned to each candidate.

We used the complex refractive indices of BC and FeOx listed in Supplementary Table 2. Ackerman and Toon49 also used this refractive index data set to estimate the radiative effects of atmospheric aerosols containing BC and magnetite. The refractive index of the coating material was assumed to 1.5+0.0i, independent of the wavelength.

Radiative transfer calculations

The spectral actinic fluxes Fac(λ) under the A-FORCE 2013W condition were calculated using the radiative transfer software package libRadtran (version 1.6 beta)50. At each altitude, the absorption coefficients babs(λ) (m−1) of the BC and FeOx particles were calculated on the basis of observational Dm/dlogDm data and theoretical σa(λ, Dm) values. The extinction coefficients bext(λ) were evaluated as babs(λ)/(1−ω), where the single-scattering albedo ω was assumed to be 0.85, independent of λ. The Henyey–Greenstein function with asymmetry parameter 0.7 was used for the scattering phase function. Shortwave absorption by gases (for example, water vapour and ozone) was calculated using the LOWTRAN/SBDART parameterization assuming the default atmospheric profiles (US-standard atmosphere). Effective solar zenith angle averaged over the daylight hours (local noon±6 h) was assumed for the actinic flux calculation. All the input parameters used in the radiative transfer calculations are listed in Supplementary Table 3.

Cloud and aerosol spectrometer probe

A cloud and aerosol spectrometer probe (CAS; Droplet Measurement Technologies, Inc.) installed under a major wing of the aircraft was used for measuring size-resolved number concentration of total aerosol in the light-scattering diameter (Dp) domain of 0.5 μm≤Dp≤50 μm. The CAS instrument used herein was also described in Koike et al.51.

Code availability

The aggregate_gen code for computing the fractal-like aggregates of spheres and the block-DDA code for electromagnetic scattering were developed by the corresponding author (N.M.) and are available in the GitHub repository at: https://github.com/nmoteki.

Data availability

FeOx data shown in Fig. 4 is available at NOAA National Center for Environmental Information (NCEI) Reference ID: YKJKWD. The other observational data and calculation results are available from N. Moteki.

Additional information

How to cite this article: Moteki, N. et al. Anthropogenic iron oxide aerosols enhance atmospheric heating. Nat. Commun. 8, 15329 doi: 10.1038/ncomms15329 (2017).

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.