id
string | text
string | image
image |
---|---|---|
170805157/2
|
et al., 2012; Larsen et al., 2016). This could reflect a real het- erogeneous distribution of $^{26}$ Al in the solar protoplanetary disc or that the dust grain carrier of $^{26}$ Al was not incorporated into chondrules (M. Bizzarro, personal communication, 2016). The abundance of $^{26}$ Al in the asteroid belt may have been as much as five times lower than the value inferred from CAIs.
Decay of $^{26}$ Al is believed to have heated some meteorite par- ent bodies to di ff erentiate completely while others (the parent bodies of the chondrites) acquired an onion layering of inter- nal heating degree below the melting temperature. The measured systematic magnetisation of some chondrites (Weiss & Elkins- Tanton,2013), supported by theoretical models of chondrule ac- cretion on existing planetesimals (Johansen et al., 2015), indi- cates that some di ff erentiated bodies have outer layers of primi- tive chondrules. A reduction in the primordial amount of $^{26}$ Al in the asteroid belt bodies, as indicated by the older Pb-Pb ages of chondrules (Connelly et al., 2012), strongly reduces the degree of internal heating and hence the expected relative frequency of fully di ff erentiated, partially di ff erentiated and primitive bodies in the asteroid belt (Larsenetal.,2016).
Chondrules appear to have crystallised after flash-heating of the unknown chondrule precursor particles (likely flu ff y dust aggregates of approximately mm sizes). An additional compo- nent of small dust grains in meteorites, the matrix, has been ar- gued to have condensed out of a solar composition gas in con- nection with chondrule formation, mainly based on the simi- lar cooling times needed to explain both chondrule and matrix crystal growth (Scott & Krot, 2005). The mechanism that led to widespread melting and partial evaporation of chondrule pre- cursors is one of the most troubling unsolved problems of me- teoritics and planet formation. Chondrule precursors must have been either heated and cooled very quickly or heated in the close vicinity of other chondrule precursors, in order to build up the saturation vapour pressure needed to stabilise the liquid rock phase, a value much in excess of the ambient pressure of the so- lar protoplanetary disc. Heating a dust aggregate in the nominal pressure of the solar protoplanetary disc will result in complete sublimation before the melting temperature is reached, unless the heating and cooling were extremely rapid. However, such rapid heating and cooling yield the wrong texture of the synthetic chondrules produced in experiments (Hewinsetal.,2005).
The requirement for chondrule precursors to heat to the liq- uid phase without vanishing by sublimation, combined with the observed lack of fractionation of volatiles in chondrules, gives a number density constraint of minimum 10 chondrule precursors per m $^3$, or a minimum mass density of approximately $10^{-5}$ kg m $^{-3}$ (Cuzzi & Alexander, 2006). Retention of the volatile ele- ment sodium during chondrule formation, inferred from the lack of internal sodium gradients within large olivine crystals, implies even higher densities, between $10^{-3}$ and $1$ kg m $^{-3}$ et al., 2008). The solar protoplanetary disc had a characteristic density of $10^{-7}$ kg m $^{-3}$ at the distance of the asteroid belt in its pristine stages and likely an order of magnitude lower after a few million years of evolution (Bitschetal.,2015). The Roche density – above which particle clumps are gravitationally bound against the tidal force from the central star – is $\rho_{\rm R} \approx 2.8
\times 10^{-5}$kg m$^{-3}$ at a distance of 2.5 AU from a solar-mass star. Chondrules therefore appear to have heated in an environment of 100 – 10,000 times the Roche density. Alexander&Ebel(2012) showed that matching the abundance of compound chondrules, fusioned dur- ing cooling at the short-lived viscous stage, also requires such ultrahigh densities. These conditions are hard to reconcile with any nominal environment of the protoplanetary disc, except per- haps very close to the star (0.1 - 0.5 AU) where the Roche density is 2 – 4 orders of magnitude higher than in the asteroid belt.
A plethora of theories have been formulated to explain the formation of chondrules. The most popular is the shock model where chondrule precursors are heated in shock waves that orig- inate in the protoplanetary disc itself or from planetesimals that plow through the gas on eccentric orbits. These models can match the right heating and cooling times (Desch & Connolly, 2002; Ciesla et al., 2004), but do not explain the high precur- sor densities needed to sustain the liquid phase and maintain volatile abundances. Current sheets driven by the magnetic field in the protoplanetary disc provide another mechanism for heat- ing chondrule precursors, but require a very high starting tem- perature of more than 800 K to reach the relevant temperature for melting chondrule precursors (McNallyetal.,2013).
The planetesimal collision model posits that chondrules are secondary in nature and arise as debris from planetesimal col- lisions. The collision model comes in two flavours: collisions between pre-melted planetesimals (Sanders & Taylor, 2005) and collisions between solid planetary embryos (Johnsonetal., 2015). The first actually taps into the decay energy of $^{26}$ Al, as the planetesimals melted through this energy input. The splash ejecta is then envisioned to crystallise as mm-sized chondrules (Asphaug et al., 2011). In the embryo model the melting hap- pens in jets that are ejected from the impact between solid em- bryos. Collisions can in principle explain the high densities of chondrule precursors as the collision debris crystallises, as well as the retention of volatiles (Dullemondetal.,2014). Producing enough chondrules to form all the chondritic bodies that must have existed in the primordial asteroid belt is nevertheless a challenge. In the model of Johnson et al. (2015), creating even 1% of the primordial asteroid belt in chondrules required a to- tal planetesimal mass in excess of three times the mass of the primordial asteroid belt, as inferred from extrapolating between the solid components of the terrestrial planets and the giant plan- ets (Hayashi, 1981). Most of the chondrules in the simulations of Johnson et al. (2015) were in fact produced in the terrestrial planet formation region where collision frequencies are higher.
Lightning as a mechanism for heating chondrule precursors was proposed by Whipple (1966). Lightning occurs in terres- trial volcanic ash plumes and causes large ash particles to melt to form chondrule-like spherules (Genareau et al., 2015). Dust aggregates hit directly by an electric discharge will nevertheless break rather than melt (G ¨ uttleretal.,2008), but exposure to the photons from the lightning results in melt spherules with size distributions similar to chondrules (Eisenhour&Buseck,1995; Poppe et al., 2010). In fact, terrestrial lightning processes ap- proximately 70% of its total energy through photons (Eisenhour etal.,1994). Radiation is attractive for chondrule melting since it naturally introduces a smallest chondrule size, in agreement with the observed log-normal or Weibull size distributions of chondrules (Dodd, 1976; Teitler et al., 2010), from the ine ffi - cient photon absorption by small dust aggregates.
Obtaining lightning discharge in protoplanetary discs re- quires (a) a charging mechanism, (b) a charge separation mech- anism and (c) low enough gas conductivity to allow the electric field to build up to a value where the current collisionally ionises the hydrogen molecule. Charging and charge separation are nat- ural consequences of vertical or radial sedimentation, since large pebbles are charged triboelectrically when colliding with dust grains (Desch & Cuzzi, 2000; Muranushi, 2010). Pilipp et al. (1998) nevertheless found that vertical sedimentation of pebbles is many orders of magnitude too slow to compete with the neu- tralising current. Desch&Cuzzi(2000) considered charge sep-
| |
170805157/11
|
Fig.5. The conditions for the nuclear battery e ff ect in the protoplanetary disc mid-plane. The topleft plot shows the pebble column density for three values of gas depletion $f_{\rm g}$ and the stopping column of positrons. The mid-plane layer transitions from optically thick to optically thin to positron stopping at a gas mass of 10% of the nominal value in the asteroid belt. The topright plot shows the breakdown value of the electric field. The bottom left plot shows the scale-height of the mid-plane layer (dotted line) together with the stopping length of the positrons at the breakdown value of the electric field. The mid-plane layer is clearly much thicker than the optimal thickness for transferring the positron energy to the electric field, but substructures of 1% of the mid-plane thickness are more in agreement with the optimal width (dot-dashed line). The bottom right plot shows the mid-plane density of gas, the Roche density and the conditions for chondrule formation. The lack of isotope fractionation and stabilisation of melted silicate can be achieved at solids-to-gas ratios of $10$ (at 1 AU) to $100$ (at 2.5 AU), but the stricter density conditions imposed by the observed high abundance of Na (Alexanderetal.,2008) can only be achieved interior of 1 AU where the Roche density is high.
## 4.2. Nuclear battery in the protoplanetary disc
The circumplanetesimal disc model presented in the previous subsection is appealing because the mid-plane pebble density reaches orders of magnitude higher than the Roche density, as required for heating chondrule precursors to the melting tem- perature and avoiding isotopic fractionation and maintaining the high abundance of Na. This model nevertheless implies that the chondrules found in chondrites were first processed in the envi- ronment near another planetesimal – and hence we must evoke a generation of planetesimals and protoplanets that formed prior to the asteroids in the asteroid belt and produced extensive amounts of chondrules that became spread throughout the protoplanetary disc.
Here we explore an alternative setting for lightning discharge and thermal processing of solids, namely the sedimented mid- plane layer in the protoplanetary disc. This mid-plane layer can not support particle densities in excess of the Roche density for long time-scales and thus the conditions are at odds with the for- mation of chondrules in the asteroid belt (we show below that the required densities can be reached inside of 1 AU). The mid-plane layer is nevertheless prone to charging by positron emission and the photons from the resulting lightning discharge could flash- heat and sublimate material in its path. These silicate vapours could in turn recondense as tiny matrix particles. Some authors have actually suggested that mineral growth in chondrules im- plies free propagation of the cooling radiation into a region of relatively low pebble density (Miura&Yamamoto,2014). Such conditions would be in agreement with the mid-plane layer of the protoplanetary disc as a chondrule formation site.
At the nominal ionisation e ffi ciency of $^{26}$ Al, we require a mass loading in small dust, $Z_1$, of at least 50 in the mid-plane layer, to suppress the neutralising current. The pebble mid-plane layer may nevertheless experience reduced positron ionisation, since the positrons streaming out from the mid-plane layer are not replaced by positrons streaming in from below and above the mid-plane layer, since those regions are underdense in pebbles. Hence the conductivity will be reduced significantly compared to thick structures. Thus even a dust-to-gas ratio as low as 10 in the mid-plane layer may yield low enough conductivities to drive charging of the mid-plane layer faster than the neutralisation.
We show in Figure 5 the conditions for the nuclear battery e ff ect in the sedimented mid-plane layer of pebbles in the proto- planetary disc. We consider a protoplanetary disc model with gas density $\rho_{\rm g} = f_{\rm
g} \times 10^{-7}\,{\rm kg\,m^{-3}}$at 2.5 AU, where$f_{\rm g}$measures the depletion of gas due to accretion, and a column density profile that falls with distance from the star$r$as$\varSigma_{\rm g} \propto r^{-1}$. The mid- plane layer scale-height is set to$H_{\rm p}=0.01 H$, a typical value
| |
170805157/6
|
density for high mass loading is then
$$
\begin{aligned}
J_{\rm dri} &=& \frac{3 \Delta v q_2}{2 \pi R_2^2 H} \approx
2.5 \times 10^{-18}\,{\rm C\,m^{-2}\,s^{-1}} \times \mathopen{}\mathclose\bgroup\originalleft(
\frac{q_2}{10^4 e} \aftergroup\egroup\originalright) \times \\
& & \mathopen{}\mathclose\bgroup\originalleft(\frac{\Delta v}{50\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright) \mathopen{}\mathclose\bgroup\originalleft(\frac{R_2}{\rm
mm} \aftergroup\egroup\originalright)^{-2} \mathopen{}\mathclose\bgroup\originalleft(\frac{r}{2.5\,{\rm AU}} \aftergroup\egroup\originalright)^{5/4} \, .
\end{aligned}
$$
The maximum radial drift current is thus more than 20 times higher than the maximum sedimentation current [equation (25)] at $Z_2 \gg 1$, but only becomes comparable to the neutralising gas current at the breakdown electric field [equation (18)] for $Z_1 \sim 10^6$.
## 2.6. Full model
The derivations presented in the previous subsections were made under a set of simplifying assumptions, e.g. that the adsorption rate of electrons and ions on dust particles are not a ff ected by the charge of the dust. We also rather handwavingly referred to the triboelectric charging model of Desch&Cuzzi(2000) when setting the characteristic charge of pebbles to $q_2
= 10^4 e$. In Appendix A and Appendix B we therefore present a fully nu- merical solution to the complex charge equation equilibrium. We include the release of secondary electrons from penetrating positrons and show that this leads to a large positive charging of the pebbles, at the expense of a negative charge on the gas. However, the results confirm that aerodynamical charge separa- tion is much too slow to compete with the neutralising current. Therefore we consider in the next section a current driven by the positrons themselves.
## 3. Nuclear battery effect
We now describe a charging mechanism akin to a nuclear bat- tery. The simplest form of a direct charging nuclear battery has a radioactive isotope on one plate of a capacitor. The nuclear de- cay products are absorbed by the opposite plate. If the emitted particles are charged (positrons, electrons, alpha particles or fis- sion fragments), then they will charge up both plates and build up a voltage di ff erence over the capacitor. In a protoplanetary disc, dense pebble structures emit positrons from the decay of $^{26}$ Al. This leads to a net negative charging of the pebbles – pro- vided that the positrons stream into regions of significant gas that stops the positrons but does not emit any positrons in the oppo- site direction. Importantly, there will only be a net current if the positrons are stopped by gas, as the dust component is in itself in complete balance between emitting and absorbing positrons.
## 3.1. Characteristic current
The characteristic positron current density arising from a pebble region, dense enough to be optically thick to its own positrons, is
$$
J_\beta = \frac{1}{6} e r_\beta \rho_{\rm p} \ell_\beta \, .
$$
Here $e$ is the positron charge, $r_\beta$ is the emission rate of positrons per unit dust mass and unit time, $\rho_{\rm p}$ is the local mass density of solids and $\ell_\beta$ is the stopping length of the positrons. We divided the expression by a factor six to take into account the random direction of the positrons. We can express the stopping length in terms of the stopping column $\varSigma_\beta = \rho_{\rm p} \ell_\beta$, yielding the positron current expression
$$
J_\beta = \frac{1}{6} e r_\beta \varSigma_\beta \approx 3.6 \times
10^{-14}\,{\rm C\,m^{-2}\,s^{-1}} \times f_{26} \, .
$$
where the activity of $^{26}$ Al relatively to the value at the for- mation conditions of CAIs, $f_{26}$, is defined as
$$
f_{26} = \frac{r_\beta}{5 \times 10^5\,{\rm kg^{-1} s^{-1}}} \, .
$$
We used here $\varSigma_\beta \approx 3$ kg / m $^2$ in silicon 1. This current density is orders of magnitude higher than the currents carried by pebble sedimentation [equation (25)] or radial drift [equation (28)] and is larger than the neutralising ion current for dust mass-loadings $Z_1 > 50$ [equation (18)]. Such mass loading of small dust in the gas is hard to achieve under nominal conditions in the pro- toplanetary disc. However, in Section 4 we apply the nuclear battery e ff ect to dense filaments in this mid-plane layer and to circumplanetesimal pebble discs. Dust is produced continuously in collisions between pebbles and hence the dust density is ex- pected to follow the density of larger pebbles. In the coagulation- fragmentation simulations of Birnstiel et al. (2011) small dust constitutes approximately 1-10% of the mass in pebbles of mm- cm sizes. We generally assume in this paper that this dust frac- tion is valid also in overdense pebble regions. The pebble-pebble collision timescale can be estimated as
$$
\begin{aligned}
t_{\rm coll} &=& (4 \pi R_2^2 n_2 \delta v)^{-1} \\
&\approx& 3 \times 10^3\,{\rm s}\, \mathopen{}\mathclose\bgroup\originalleft(\frac{R_2}{\rm mm} \aftergroup\egroup\originalright) \mathopen{}\mathclose\bgroup\originalleft(
\frac{\rho_2}{10^{-2}\,{\rm kg\,m^{-3}}} \aftergroup\egroup\originalright)^{-1} \mathopen{}\mathclose\bgroup\originalleft(\frac{\delta
v}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright)^{-1} \, .
\end{aligned}
$$
The first lightning discharge will thus be delayed until small dust has been produced to reduce the conductivity. However, the dust production time-scale to produce small dust fragments is much shorter than the decay time of $^{26}$ Al, within the range of likely physical parameters in equation (32). We assume that the dust within dense pebble filaments undergoes continuous release in pebble-pebble collisions and recoagulation to maintain a high dust mass loading for millions of years.
As we discuss further in the next subsection, the neutralising gas current in dense pebble structures may be even lower than the expression in equation (18), if the gradient in the particle density is high. In that case positrons streaming out of the region are not replaced by positrons streaming in and hence the ionisation rate inside of the structure is reduced.
## 3.2. Optimal conditions for the nuclear battery
Consider a dense pebble region emitting positrons that are ab- sorbed in the surrounding gas. The charging of the pebbles builds up an electric field that points from the surroundings into the negatively charged pebble region. Initially the positrons are stopped by collisions with gas molecules, but as the pebble re- gion becomes increasingly negatively charged, the positrons lose more and more of their kinetic energy to work against the elec- tric field. The width of the pebble structure is now key to ob- taining neutralisation through lightning discharge. If the pebble structure has a too large width, $W$, then the positrons are trapped in the pebble region before the breakdown electric field value is reached. The critical width $W^\dagger$ is given through the relation
$$
T_\beta = e W^\dagger E_{\rm BD} \, .
$$
1 We calculated the stopping length $\ell_\beta \approx 1.3$ mm for 0.66 MeV positrons and internal density $\rho_\bullet=2.3 \times 10^3$ kg m $^{-3}$, using the positron penetration calculations presented in Appendix D; see also Umebayashi et al. (2013) who obtained a similar value of the positron stopping length.
| |
170805157/4
|
denote the e ff ective radii and masses that are available for ab- sorbing and releasing positrons. These quantities will only be di ff erent from $R_1$, $R_2$, $m_1$ and $m_2$ if the particles are larger than $\chi_{\rm
s}$. The e ff ective radius can be written$R_i^*={\rm min}(R_i,\chi_{\rm s})$and the e ff ective mass$m_i^* = m_i - m_i (R<\chi_{\rm s})$, where we subtract the inactive mass present deeper than$\chi_{\rm s}$ in the solid.
We can calculate the positron production rate per kilogram of rocky material from the expression
$$
r_\beta = \frac{b x_{26}}{\tau m_{\rm H_2} Z_{\rm sil}} \, .
$$
Here $b=0.82$ is the branching ratio for positron emission, $x_{26}=3\times10^{-10}$ is the nominal ratio of $^{26}$ Al nuclei to hydrogen molecules, $\tau$ is the decay time of $^{26}$ Al (1.03 Myr) and $Z_{\rm sil}$ is the dust mass fraction in rock (we use the approximate solar com- position value of 0.005, following Lodders, 2003). This gives $r_\beta= 5
\times 10^{5} \, {\rm kg^{-1} s^{-1}}$.
In regions of the protoplanetary disc where the number den- sity of dust is very high, the number densities of ions and elec- trons are given by a balance between ionisation, at rate $\zeta$ per neutral molecule, and absorption by dust grains (e.g.Okuzumi, 2009),
$$
\begin{aligned}
n_{\rm i} &=& \frac{\zeta n_{\rm n}}{n_d \sigma_d v_{\rm i}} \, ,
\\
n_{\rm e} &=& \frac{\zeta n_{\rm n}}{n_d \sigma_d v_{\rm e}} \,
.
\end{aligned}
$$
$$
\begin{aligned}
n_{\rm i} &=& \frac{\zeta n_{\rm n}}{n_d \sigma_d v_{\rm i}} \, ,
\\
n_{\rm e} &=& \frac{\zeta n_{\rm n}}{n_d \sigma_d v_{\rm e}} \,
.
\end{aligned}
$$
The index $d$ introduced here implies summation over dust species. The ion and electron speeds, $v_{\rm i}$ and $v_{\rm e}$, are discussed fur- ther in Section 2.3. The ionisation rate by positrons, $\zeta$, is given by the expression
$$
\zeta = n_\beta \langle \sigma_{\rm n\beta} v_\beta \rangle \, .
$$
Here $\sigma_{\rm n\beta}$ is the ionisation cross section of molecular hydro- gen to positrons
$$
\sigma_{\rm n\beta} = \frac{m_{\rm n}}{\varSigma_\beta} \frac{T_\beta}{E_{\rm
ion}} \approx 2.0 \times 10^{-23}\,{\rm m^2} \, ,
$$
where we have approximated the stopping column density in molecular hydrogen by that in silicon, $\varSigma_\beta$. The number density $n_\beta$ of positrons in the gas, moving at the speed $v_\beta$, follows directly from the production equation
$$
\dot{n}_\beta = r_\beta \rho_{\rm p}^* - n_\beta n_d \sigma_d v_\beta
(R_d^*/\chi_{\rm s}) - n_\beta n_{\rm n} \langle \sigma_{\rm n\beta} v_\beta
\rangle (E_{\rm ion}/T_\beta) \, .
$$
The first term represents the emission rate, with $\rho_{\rm p}^* = n_1 m_1^* +
n_2 m_2^*$. The second term represents the absorption in solid par- ticles and the third term the absorption in gas molecules. The equilibrium state with$\dot{n}_\beta=0$ yields
$$
n_\beta = \frac{r_\beta \rho_{\rm p}^*}{n_d \sigma_d v_\beta
(R_d^*/\chi_{\rm s}) + n_{\rm n} \langle \sigma_{\rm n\beta} v_\beta
\rangle (E_{\rm ion}/T_\beta)} \, .
$$
We now rewrite the denominator in terms of the stopping column density $\varSigma_\beta = \rho_\bullet \chi_{\rm s} = \rho_{\rm g} \chi_{\rm s,g}$, where we introduced the material density $\rho_\bullet$ and the stopping length in the gas $\chi_{\rm s,g}=(n_{\rm n} \sigma_{\rm n \beta})^{-1} (T_\beta/E_{\rm
ion})$and assumed that the stopping column is in- dependent of the material (molecular hydrogen or silicate rock). That yields now, for solid particles with radii much smaller than$\chi_{\rm s}$, the approximate expression
$$
n_\beta \approx \frac{r_\beta \rho_{\rm p}}{\rho_{\rm p}
v_\beta/\varSigma_\beta + \rho_{\rm g} v_\beta/\varSigma_\beta} \, .
$$
Combining equations (9) and (5) yields
$$
\zeta = \zeta_0 \frac{Z/0.005}{1+Z}
$$
with $\zeta_0=0.005 r_\beta \varSigma_\beta \sigma_{{\rm n}\beta} \approx 1.5
\times 10^{-19}\,{\rm s^{-1}}$denoting the ionisation rate at the unsedimented value$Z=0.005$. Here$Z=Z_1+Z_2$is the total mass loading of solid particles in the gas. The ionisation rate thus rises from$\zeta_0$for unsedimented par- ticles well-mixed with the gas in the entire vertical column to$\zeta_\infty
= 200 \zeta_0 \approx 3 \times 10^{-17}\,{\rm s^{-1}}$when the particle mass loading is high,$Z\gg1$, e.g. in a dense mid-plane layer of sedimented peb- bles.
## 2.3. Ion and electron currents
Knowing the ion and electric densities from equation (3) and (4) allows us to calculate the neutralising current of ions and electrons as a function of the electric field strength $E$ through
$$
\begin{aligned}
{\boldsymbol J}_{\rm i} &=& -e n_{\rm i} \langle {\boldsymbol v}_{\rm i} \rangle \, ,
\\
{\boldsymbol J}_{\rm e} &=& +e n_{\rm e} \langle {\boldsymbol v}_{\rm e} \rangle
\, .
\end{aligned}
$$
$$
\begin{aligned}
{\boldsymbol J}_{\rm i} &=& -e n_{\rm i} \langle {\boldsymbol v}_{\rm i} \rangle \, ,
\\
{\boldsymbol J}_{\rm e} &=& +e n_{\rm e} \langle {\boldsymbol v}_{\rm e} \rangle
\, .
\end{aligned}
$$
The mean ion and electron velocities, $\langle {\boldsymbol v}_{\rm i} \rangle$ and $\langle {\boldsymbol v}_{\rm e} \rangle$ follow from the expression (Mori&Okuzumi,2016)
$$
\langle {\boldsymbol v}_\alpha \rangle = \frac{m_\alpha + m_{\rm n}}{m_\alpha m_{\rm
n}} e {\boldsymbol E} \Delta t_\alpha \, .
$$
Here $\alpha$ denotes either ions or electrons. The mean time be- tween collisions is given for electrons as
$$
\Delta t_{\rm e} = \frac{1}{n_{\rm n} \langle \sigma_{\rm en} v_{\rm e}
\rangle}
$$
and for ions as
$$
\Delta t_{\rm i} = \frac{1}{1.6 \times 10^{-15}\,{\rm m^3\,s^{-1}} \times
n_{\rm n}} \, .
$$
The constant in the denominator expresses that the interac- tion cross section between ions and neutrals actually decreases with increasing ion speed, yielding a constant product of cross section and speed (Wannier, 1953). The current densities in equations (11) and (12) are proportional to $E$ (through equation 13) only if $n_{\rm i}$ and $n_{\rm e}$ are independent of $E$. In fact, as first pointed out by Okuzumi&Inutsuka(2015), $n_{\rm i}$ and $n_{\rm e}$ can decrease with increasing $E$ when the electric field is su ffi ciently strong. If $E$ is above a threshold [see equations (19) and (20) for the thresholds for ion and electron heating, respectively], the electrons and ions obtain kinetic energies above the mean thermal energy of the neutrals through acceleration by the electric field. In the pres- ence of such a strong electric field, the plasma particles collide with and adsorb onto dust grains more frequently as the field strength increases, and hence their number density decreases with increasing $E$. For the ion component, the added energy re- sides mainly in the mean motion because of the larger inertia of the ion molecules (e.g., ${\rm HCO^+}$) than that of neutrals (mainly ${\rm H_2}$ and He). Hence, for electrically heated ions, the magnitude of the mean velocity $\langle {\boldsymbol v}_{\rm i} \rangle$ in equation (11) becomes equal to the mean speed $\langle v_{\rm i} \rangle$ in equation (3). Using this fact together with equations (3) and (11), one can show that the neutralising ion current ap- proaches the constant expression (Okuzumi & Inutsuka, 2015)
$$
J_{\rm i,\infty} = \frac{e \zeta n_{\rm n}}{n_d \sigma_d} \, .
$$
| |
170805157/10
|
Fig.4. The conditions for the nuclear battery e ff ect in circumplanetesimal discs. The topleft plot shows the pebble column density as a function of the distance from the planetesimal, for two values of the planetesimal radius ($R_{\rm p}=100\,{\rm km}$ and $R_{\rm p}=1000\,{\rm km}$) and three values of the random pebble speed $c$. The vertical dot-dashed lines indicate the Hill radius at 2.5 AU and 25 AU, respectively, and the horizontal dashed line indicates the stopping column of positrons. The pebble column density is below the stopping column of positrons in the outer region of the circumplanetesimal disc, unless the random pebble motion is very high, so positron emission leads to an e ffi cient charging of the circumplanetesimal disc. The top right plot shows the integrated disc mass relative to the mass of the central planetesimal. The small planetesimal has a relatively more massive disc than the larger protoplanet and hence can convert a higher fraction of the central mass to chondrules. The bottomleft plot shows the scale- height of the pebble disc. The stopping length of positrons at the breakdown value of the electric field is indicated for the nominal breakdown current at 2.5 AU. The bottomright plot shows the mid-plane density of pebbles compared to the minimum pebble density required for lack of isotopic fractionation and stability of the liquid phase by Cuzzi&Alexander(2006) and the density range implied by the observed high abundance of Na within chondrules by Alexanderetal.(2008). The pebble density in the circumplanetesimal disc environment lies perfectly within these experimental constraints.
trates well the need for computer simulations to understand the formation and dynamics of circumplanetesimal discs better.
Pebbles in the circumplanetesimal disc will spend most of their time in the outer regions of the disc where the accretion time-scale is longest. The viscous time-scale of a circumplan- etesimal particle disc of optical depth $\tau$ (= $\varSigma_{\rm p} \sigma_{\rm p}/m_{\rm p}$) is (seee.g.
$$
T = \mathopen{}\mathclose\bgroup\originalleft(\frac{r}{c} \aftergroup\egroup\originalright)^2 \varOmega_{\rm pla} \frac{1+\tau^2}{\tau}
$$
The viscous time-scale in the limit $\tau \gg 1$ can now be ex- pressed as
$$
\begin{aligned}
T &=& \mathopen{}\mathclose\bgroup\originalleft(\frac{r}{c} \aftergroup\egroup\originalright)^2 \varOmega_{\rm pla} \frac{c \varOmega_{\rm
pla}}{\pi G} \frac{\pi R_{\rm p}^2}{m_{\rm p}} = \frac{R_{\rm pla}^3}{c r
R_{\rm p}} \\
&=& 3.2 \,{\rm Myr}\, \mathopen{}\mathclose\bgroup\originalleft(\frac{r}{100\,R_{\rm pla}} \aftergroup\egroup\originalright)^{-1} \mathopen{}\mathclose\bgroup\originalleft(
\frac{c}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright)^{-1} \mathopen{}\mathclose\bgroup\originalleft(\frac{R_{\rm
pla}}{10^3\,{\rm km}} \aftergroup\egroup\originalright)^2 \mathopen{}\mathclose\bgroup\originalleft(\frac{R_{\rm p}}{1\,{\rm mm}}
\aftergroup\egroup\originalright)^{-1} \, .
\end{aligned}
$$
This is clearly a very long time-scale for a protoplanet of ra- dius $R_{\rm
p}=10^3\,{\rm km}$, which validates our argument that accretion will happen by gravitational stresses and thus will be regulated by the continuous infall of new pebble material onto the circum- planetesimal disc. A planetesimal of radius$R_{\rm p}=100\,{\rm km}$has a much shorter accretion time-scale, unless the random pebble motion is less than$0.1\,{\rm m\,s^{-1}}$.
Johansen et al. (2015) showed that planetesimals can grow to protoplanets of Mars mass, initially by planetesimal- planetesimal collisions and later mainly by pebble accretion, in the asteroid belt after a few million years. The pebble accre- tion process would likely endow these protoplanets with exten- sive circumplanetesimal pebble discs. Drazkowskaetal.(2016) demonstrated that the pile up of particles in the terrestrial planet region can trigger planetesimal formation already after a few hundred thousand years evolution of the protoplanetary disc. Therefore protoplanets may have populated the terrestrial planet region even earlier than the asteroid belt region and hence chon- drules could have been produced around these protoplanets and subsequently transported to the asteroid formation region to become incorporated into the asteroids and protoplanets that formed there.
| |
170805157/13
|
taken to follow the so-called Townsend scalings (Ebert et al., 2010), which yield a discharge length-scale proportional to the number density of neutrals $n$, a discharge time-scale proportional to $1/n$ and a discharge current independent of $n$. Hood&Kring (1996) used the 100 ms duration of terrestrial lightning to argue that protoplanetary disc lightning has a characteristic time-scale of $10^4$ – $10^5$ s. However, terrestrial lightning strokes consist of several flashes of $30$
$\mu$ s
duration, which would yield $10^2$ s time- scales in the asteroid belt region of the hydrogen-dominated so- lar protoplanetary discs with $\rho_{\rm g}=10^{-7}\,{\rm
kg\,m^{-3}}$.
Note that the central temperature in equation (55) has a very weak scaling with $\tau_\star$ – even a dramatic increase in the optical depth will not lead to a significant temperature increase in the core of the filament. This is a result of the radial thinning of the radiation at large distances from the lightning. The temperature at the optical at the optical surface ($\tau=1/2$) relates to the deep temperature $T_\star$ through
$$
T(1/2) = \frac{T_\star}{[\tau_\star \ln (2 \tau_\star)]^{1/4}} \, .
$$
That gives $T(1/2)=366$ K for $T_\star=1754$ K and $\tau_\star=100$. The total energy content of the pebble region at the equilibrium temperature is obtained by integrating over the energy contents of the cylindrical annuli,
$$
Q = \int_0^{R_\star} c_{\rm p} \rho_{\rm p} T 2 \pi R h \mathrm{d} R =
2 \pi h c_{\rm p} \rho_{\rm p} (\Delta R)^2 \int_{0}^{\tau_*} T \tau \mathrm{d}
\tau \, .
$$
Inserting equation (54) we obtain the expression
$$
\begin{aligned}
Q &=& \mathopen{}\mathclose\bgroup\originalleft(\frac{I E}{2 \pi \sigma_{\rm SB} \Delta R} \aftergroup\egroup\originalright)^{1/4} 2 \pi h
c_{\rm p} \rho_{\rm p} (\Delta R)^2
\\ & & \times \int_0^{\tau_*} \mathopen{}\mathclose\bgroup\originalleft[\frac{1}{\tau_\star} + \ln
(\tau_\star-1/2) - \ln(\tau_\star-\tau) \aftergroup\egroup\originalright]^{1/4} \tau \mathrm{d} \tau \, .
\end{aligned}
$$
We find empirically that the term in hard brackets within the in- tegral has a mean value, weighted by $\tau$, of approximately unity for a wide range of values of $\tau_\star$. Hence the integral is well ap- proximated by $(1/2)
\tau_\star^2$. That gives the energy content per length
$$
\begin{aligned}
Q/h &=& 6.5 \times 10^8\,{\rm J\,m^{-1}} \times \mathopen{}\mathclose\bgroup\originalleft(\frac{\tau_\star}{100}
\aftergroup\egroup\originalright)^2 \\ & & \times \mathopen{}\mathclose\bgroup\originalleft(
\frac{I}{10^5\,{\rm A}} \aftergroup\egroup\originalright)^{1/4} \mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_{\rm g}}{10^{-7}\,{\rm
kg\,m^{-3}}} \aftergroup\egroup\originalright)^{1/4} \\
& & \times \mathopen{}\mathclose\bgroup\originalleft(\frac{c_{\rm p}}{10^3\,{\rm J\,kg^{-1}\,K^{-1}}} \aftergroup\egroup\originalright)
\mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_2}{10^{-2}\,{\rm kg\,m^{-3}}} \aftergroup\egroup\originalright)
\mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_1}{10^{-3}\,{\rm kg\,m^{-3}}} \aftergroup\egroup\originalright)^{-7/4}
\\ & & \times \mathopen{}\mathclose\bgroup\originalleft(\frac{R_1}{10^{-6}\,{\rm m}} \aftergroup\egroup\originalright)^{7/4} \mathopen{}\mathclose\bgroup\originalleft(
\frac{\rho_\bullet}{10^{3}\,{\rm kg\,m^{{-3}}}} \aftergroup\egroup\originalright)^{7/4} \, .
\end{aligned}
$$
Here we assumed that the density of solid particles is dominated by the pebble component, $\rho_{\rm p} \approx \rho_2$. The time to heat the dense pebble region is
$$
\begin{aligned}
t_{\rm heat} &=& \frac{Q/h}{I E} = 7.7 \times 10^2\,{\rm s}\, \mathopen{}\mathclose\bgroup\originalleft(
\frac{\tau_\star}{100} \aftergroup\egroup\originalright)^2 \\
& & \times \mathopen{}\mathclose\bgroup\originalleft(
\frac{I}{10^5\,{\rm A}} \aftergroup\egroup\originalright)^{-3/4} \mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_{\rm g}}{10^{-7}\,{\rm
kg\,m^{-3}}} \aftergroup\egroup\originalright)^{-3/4} \\
& & \times \mathopen{}\mathclose\bgroup\originalleft(\frac{c_{\rm p}}{10^3\,{\rm J\,kg^{-1}\,K^{-1}}} \aftergroup\egroup\originalright)
\mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_2}{10^{-2}\,{\rm kg\,m^{-3}}} \aftergroup\egroup\originalright)
\mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_1}{10^{-3}\,{\rm kg\,m^{-3}}} \aftergroup\egroup\originalright)^{-7/4}
\\
& & \times
\mathopen{}\mathclose\bgroup\originalleft(\frac{R_1}{10^{-6}\,{\rm m}} \aftergroup\egroup\originalright)^{7/4}
\mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_\bullet}{10^{3}\,{\rm kg\,m^{{-3}}}} \aftergroup\egroup\originalright)^{7/4} \, .
\end{aligned}
$$
Fig.6. The temperature as a function of distance from the centre of the lightning discharge channel for a current $I=10^5\,{\rm A}$, gas density $\rho_{\rm g} = 3 \times 10^{-7}\,{\rm kg\,m^{-3}}$, dust density $\rho_1=10^{-3}\,{\rm kg\,m^{-3}}$, pebble density $\rho_2=10^{-2}\,{\rm
kg\,m^{-3}}$and optical depth$\tau_\star=300$. The discharge channel has a width of 100 m, indicated by the vertical dotted line. The discharge is assumed to last$2 \times 10^3$s, followed by a cooling phase that is evolved up to a time of$10^4$ s. Heating is rapid and uniform within the discharge channel, while the outer regions only heat after radiation di ff uses out from the discharge region. The equilibrium temperature follows the an- alytical solution outside of the discharge channel.
## 5.2. Heating and cooling rates
Chondrule precursors appear to have been heated to near- liquidus temperatures, reaching the range 1500 K – 2200 K (Hewins&Radomsky,1990). This constraint is mainly based on what appears to be relict, unmelted grains inside chondrules. If these grains are instead interpreted as interlopers that penetrated a fully molten spherule (Connolly&Hewins,1995), then chon- drule precursors could have been heated to much higher tem- peratures. The cooling rates of chondrules can be constrained by studying crystal growth in furnace experiments of chondrule analogues or from the retention of volatiles such as Na in chon- drules. Hewins&Radomsky(1990) found that cooling rates in the range $10^2$ – $10^3$ K / hr are consistent with experimental data. The upper range of these cooling rates corresponds to what is in- ferred for matrix grains (Scott&Krot,2005). On the other hand, Yurimoto&Wasson(2002) and Miura&Yamamoto(2014) in- ferred much higher rates, in the range $10^5$ – $10^7$ K / hr. The upper range of those rates are comparable to cooling rate of isolated molten droplets,
$$
\begin{aligned}
\dot{T} &=& \frac{4 \pi R_2^2 \sigma_{\rm SB} T^4}{c_{\rm p} (4 \pi/3)
\rho_\bullet R_2^3} \\ &\approx& 9.8 \times 10^6\,{\rm
K\,s^{-1}}\,\mathopen{}\mathclose\bgroup\originalleft(\frac{T}{2000\,{\rm K}} \aftergroup\egroup\originalright)^4 \mathopen{}\mathclose\bgroup\originalleft(
\frac{\rho_\bullet}{10^3\,{\rm kg\,m^{-3}}} \aftergroup\egroup\originalright)^{-1} \mathopen{}\mathclose\bgroup\originalleft(
\frac{R_2}{1\,{\rm mm}} \aftergroup\egroup\originalright)^{-1} \, .
\end{aligned}
$$
Much lower cooling rates can be achieved in an optically thick region of chondrules. We measure the heating and cooling rates of pebbles exposed to radiation from lightning discharge by solv- ing equation (48) numerically starting at a background temper- ature of $T=200$ K. Absorption and release of latent heat $\mathcal{L}$ by melting and crystallisation is included using a simplified ap- proach where we translate the internal energy $e$ to temperature
| |
170805157/21
|
For the repulsion case with $\phi_d > 0$ we obtain
$$
\begin{aligned}
K_{d{\rm i}} &=& \pi R_d^2\Biggl\{ \sqrt{\frac{k_{\rm B}T}{2\pi m_{\rm i}}}
\mathopen{}\mathclose\bgroup\originalleft[\mathopen{}\mathclose\bgroup\originalleft(1+\frac{v_0}{|\langle\bm v_{\rm
i}\rangle|}\aftergroup\egroup\originalright)\exp\mathopen{}\mathclose\bgroup\originalleft(-\frac{m_{\rm i}(v_0-|\langle\bm v_{\rm
i}\rangle|)^2}{2k_{\rm B}T_{\rm i}}\aftergroup\egroup\originalright) + \aftergroup\egroup\originalright. \\
&& \mathopen{}\mathclose\bgroup\originalleft. \mathopen{}\mathclose\bgroup\originalleft(1-\frac{v_0}{|\langle\bm v_{\rm
i}\rangle|}\aftergroup\egroup\originalright)\exp\mathopen{}\mathclose\bgroup\originalleft(-\frac{m_{\rm i}(v_0+|\langle\bm v_{\rm
i}\rangle|)^2}{2k_{\rm B}T_{\rm i}}\aftergroup\egroup\originalright) \aftergroup\egroup\originalright] \Biggr. + \\
&& \Biggl. \frac{|\langle\bm v_{\rm i}\rangle|}{2}\mathopen{}\mathclose\bgroup\originalleft(1+\frac{k_{\rm B}T_{\rm
i}-2e\phi_d}{m_{\rm i} |\langle\bm v_{\rm i}\rangle|^2}\aftergroup\egroup\originalright) \mathopen{}\mathclose\bgroup\originalleft[{\rm
erf}\mathopen{}\mathclose\bgroup\originalleft(\sqrt{\frac{m_{\rm i}}{2k_{\rm B}T_{\rm i}}}(|\langle\bm v_{\rm
i}\rangle|-v_0)\aftergroup\egroup\originalright) + \aftergroup\egroup\originalright. \\
&& \mathopen{}\mathclose\bgroup\originalleft. {\rm erf}\mathopen{}\mathclose\bgroup\originalleft(\sqrt{\frac{m_{\rm i}}{2k_{\rm B}T_{\rm i}}}(|\langle\bm v_{\rm
i}\rangle|+v_0)\aftergroup\egroup\originalright) \aftergroup\egroup\originalright] \,\, \Biggr\} \, .
\end{aligned}
$$
For the attraction case with $\phi_d < 0$ we obtain an expression already given in Shukla&Mamun(2002),
$$
\begin{aligned}
K_{d{\rm i}} &=& \pi R_d^2 \mathopen{}\mathclose\bgroup\originalleft\{\sqrt{\frac{2k_{\rm B}T}{\pi m_{\rm i}}}
\exp\mathopen{}\mathclose\bgroup\originalleft(-\frac{m_{\rm i} |\langle\bm v_{\rm i}\rangle|^2}{2k_{\rm B}T_{\rm i}}
\aftergroup\egroup\originalright) \aftergroup\egroup\originalright. \\
&& \mathopen{}\mathclose\bgroup\originalleft. + |\langle\bm v_{\rm i}\rangle|\mathopen{}\mathclose\bgroup\originalleft(1+\frac{k_{\rm B}T_{\rm
i}-2e\phi_d}{m_{\rm i} |\langle\bm v_{\rm i}\rangle|^2}\aftergroup\egroup\originalright) {\rm
erf}\mathopen{}\mathclose\bgroup\originalleft(\sqrt{\frac{m_{\rm i}}{2k_{\rm B}T_{\rm i}}}|\langle\bm v_{\rm
i}\rangle|\aftergroup\egroup\originalright) \aftergroup\egroup\originalright \} \, .
\end{aligned}
$$
Note that the repulsion and accretion expressions for $K_{d{\rm i}}$ are the same for $\phi_d=v_0=0$.
## C.1. Dust adsorption
The adsorption of small dust grains by large pebbles is given by the expression
$$
K_{12} = \sigma_2 v_{12} \mathopen{}\mathclose\bgroup\originalleft(1 - \frac{q_1 e \phi_2}{E_1} \aftergroup\egroup\originalright) \, .
$$
Here $v_{12}$ is the relative drift speed between dust grains and pebbles, calculated from the drag force equilibrium of Equations B.5–B.10, and $E_1$ is the kinetic energy
$$
E_1 = \frac{1}{2} m_1 v_{12}^2 \, .
$$
## AppendixD: Secondary electrons
Positrons released by $^{26}$ Al in regions of very high particle den- sity are primarily stopped by other particles. The passage of a positron through the surface of a solid particle is accompanied by the emission of $N_{\rm s}$ secondary electrons. This leads to a net negative charging of dust particles, since a small fraction of the positrons are absorbed by the gas. The secondary electrons also contribute to the conductivity of the gas. In this appendix we cal- culate the yield of secondary electrons as a function of the size of the particle that the positron passes through. The secondary electron yield is an important parameter in the charge equilib- rium model presented in the main paper.
The production of secondary electrons by a positron expe- riencing energy loss $\mathrm{d} E_\beta$ in the interval between $\chi$ and $\chi+\mathrm{d}\chi$ along its penetration path through a solid particle can be written according to Lin&Joy(2005) as
$$
\mathrm{d} N_{\rm sec} = - \frac{\mathrm{d} E_\beta}{\mathrm{d} \chi} \frac{1}{\epsilon} 0.5
\exp(-\alpha \chi_\perp) \mathrm{d} \chi \, .
$$
Here $\epsilon$ is the ionisation potential of electrons inside the mate- rial, $\alpha$ is the inverse stopping length of the released electrons and $\chi_\perp$ is the shortest distance to the surface of the particle. The par- ticle sizes we consider are much larger than the stopping length
Fig.D.1. Secondary electron production versus the incident angle $\mu=\cos(\theta)$ for particles of radii $\mu$ m, mm and cm. The secondary electron production rises with decreasing $\mu$ and experiences a jump at the angle when the positron penetrates all the way through the solid. Finally, the secondary electron production falls when the pathway through the par- ticle is short at very low $\mu$ (dotted line). The integrated yield (dashed lines) is about 0.08 for mm and cm solids and 0.04 for $\mu$ m solids.
of the electrons and hence the value of $\alpha$ obtained from exper- iments with flat surfaces can be applied to curved particles as well. The perpendicular distance $\chi_\perp$ can be written as function of $\chi$ for a positron with angle of incidence $\theta$ relative to the nor- mal of the surface as
$$
\chi_\perp = \sqrt{R^2 + \chi^2 - 2 R \chi \cos(\theta)} \, .
$$
We use the parameters $\epsilon=90$ eV and $\lambda=1/\alpha=2.7$ nm rel- evant for pure silicon (Lin&Joy,2005). The energy loss rate is calculated from the Bethe-Bloch equation modified for positrons (e.g.Oliveetal.,2014,chapter31),
$$
\frac{\mathrm{d} E}{\mathrm{d} \chi} = -\frac{1}{2} K \rho_{\rm i} \frac{Z_{\rm i}}{A_{\rm
i}} \frac{1}{\beta^2} \mathopen{}\mathclose\bgroup\originalleft\{\ln \frac{m_{\rm e} c^2 \beta^2 \gamma^2
[m_{\rm e} c^2 (\gamma-1)]}{2 I_{\rm i}^2} + F \aftergroup\egroup\originalright\} \, .
$$
Here the constant $K=4 \pi N_{\rm A}r_{\rm e}^2 m_{\rm e} c^2$, with $N_{\rm
A}$denoting Avogadro’s number,$r_{\rm e}$the classical electron radius,$m_{\rm e}$the elec- tron mass and$c$the speed of light. The material is characterised by the density$\rho_{\rm i}$, nuclear charge$Z_{\rm i}$and atomic number$A_{\rm i}$. The positron energy is defined through$\beta = v/c$and$\gamma = 1/\sqrt{1-\beta^2}$. The function$F$ is specific to positrons and is defined as
$$
F = 2 \ln 2 - \frac{\beta^2}{12} \mathopen{}\mathclose\bgroup\originalleft[23 + \frac{14}{\gamma+1} +
\frac{10}{(\gamma+1)^2} + \frac{4}{(\gamma+1)^3} \aftergroup\egroup\originalright] \, .
$$
We calculate the number of secondary electrons produced as a function of $\mu =
\cos(\theta)$for particles of$\mu$m, mm and cm sizes. Secondary electrons are mainly produced within a few times the mean free path$\lambda$from the surface. We therefore only integrate Eq. (D.1) at distances within$10
\lambda$from the surface, in order to speed up the integration. We use an energy of$E_0=0.66$MeV for the positron. The results are shown in Fig. D.1. The yield of secondary electrons increases initially with decreasing$\mu$, as
| |
170805157/22
|
the positron moves closer to the surface. The yield experiences a sudden increase at an angle where the positron passes all the way through the particle. Finally, the yield falls along a characteristic curve
$$
N_{\rm sec}=\frac{2 R \mu}{\chi_{\rm s}} \frac{0.5 E_0}{\epsilon}
$$
for very low values of $\mu$. Here $2 R \mu$ is the passage length through the particle and $\chi_{\rm s}$ is the total stopping length in the material (1 mm for silicate particles). The integrated yield gives the total number of secondary electrons through
$$
N_{\rm sec} = (2 \pi)^{-1} \int_0^{2 \pi} \int_0^{\pi/2} \mathrm{d} N_{\rm sec}
\sin \theta \mathrm{d} \theta \mathrm{d} \phi = \int_0^1 \mathrm{d} N_{\rm sec} \mathrm{d} \mu \, .
$$
This gives an integrated yield of $N_{\rm sec}=0.08$ for particles of mm and cm sizes and $N_{\rm sec}=0.04$ for $\mu$ m-sized particles.
| |
170805157/1
|
# Harvesting the decay energy of $^{26}$ Al to drive lightning discharge in protoplanetary discs
Anders Johansen 1 & Satoshi Okuzumi
Chondrules in primitive meteorites likely formed by recrystallisation of dust aggregates that were flash-heated to nearly complete melting. Chondrules may represent the building blocks of rocky planetesimals and protoplanets in the inner regions of protoplanetary discs, but the source of ubiquitous thermal processing of their dust aggregate precursors remains elusive. Here we demonstrate that escape of positrons released in the decay of the short-lived radionuclide $^{26}$ Al leads to a large-scale charging of dense pebble structures, resulting in neutralisation by lightning discharge and flash-heating of dust and pebbles. This charging mechanism is similar to a nuclear battery where a radioactive source charges a capacitor. We show that the nuclear battery e ff ect operates in circumplanetesimal pebble discs. The extremely high pebble densities in such discs are consistent with conditions during chondrule heating inferred from the high abundance of sodium within chondrules. The sedimented mid-plane layer of the protoplanetary disc may also be prone to charging by the emission of positrons, if the mass density of small dust there is at least an order of magnitude above the gas density. Our results imply that the decay energy of $^{26}$ Al can be harvested to drive intense lightning activity in protoplanetary discs. The total energy stored in positron emission is comparable to the energy needed to melt all solids in the protoplanetary disc. The e ffi ciency of transferring the positron energy to the electric field nevertheless depends on the relatively unknown distribution and scale-dependence of pebble density gradients in circumplanetesimal pebble discs and in the protoplanetary disc mid-plane layer.
Keywords. meteorites, meteors, meteoroids – minor planets, asteroids: general – planets and satellites: formation – protoplanetary disks
## 1. Introduction
Tiny spherules that crystallised from molten rock are abundant in primitive chondrite meteorites from the asteroid belt (Krotet al.,2009). These chondrules of typical sizes between 0.1 and 1 mm comprise approximately 60-80% of the mass of the ordinary and enstatite chondrite classes, mixed with smaller metal grains and microscopic matrix particles (Dodd, 1976). The chondrule abundance is more varied in the carbonaceous chondrites, rang- ing from approximately 50% in CO and CV chondrites to 0% in the CI chondrites (Scott & Krot, 2003). The lack of chon- drules in the CI chondrites could nevertheless be the result of chondrule destruction during extensive aqueous alteration of the parent body (Endressetal.,1996).
The sheer abundance of chondrules implies that these parti- cles played a key role in the formation and / or growth of planetes- imals in the solar protoplanetary disc. Chondrule-sized particles can concentrate into dense clumps in the gaseous protoplanetary disc either by preferential concentration between turbulent ed- dies near the dissipation scale of the turbulent gas (Cuzzietal., 2008) or by the streaming instability that concentrates particles into dense filaments through their collective drag force on the gas (Youdin&Goodman,2005;Johansen&Youdin,2007;Bai & Stone, 2010a). The streaming instability concentrates parti- cles as small as chondrules at the location of the asteroid belt if the turbulence in the protoplanetary disc is su ffi ciently weak to allow mm-sized particles to sediment (Carreraetal.,2015;Yang &Johansen,2017). Chondrules also have ideal sizes to become accreted onto young planetesimals and protoplanets. Johansen etal.(2015) showed that the observed size distribution of aster- oids (particularly the steeply declining di ff erential size distribu- tion for asteroids larger than 100 km in diameter and the transi- tion to a shallower size distribution above 400 km in diameter) can arise due to the accretion of chondrules with sizes similar to those measured in ordinary chondrites. Chondrule accretion also drives the continued growth to Mars-sized protoplanets, both in the asteroid belt and in the terrestrial planet region. Therefore understanding the origin of chondrules is imperative for under- standing the formation of asteroids and terrestrial planets.
The identification of radiogenic $^{26}$ Mg in the CAI compo- nent (calcium-aluminium-rich inclusions) of chondrites, pro- duced through the decay of the short-lived radionuclide $^{26}$ Al (Lee et al., 1977), provides a strong connection between the solar protoplanetary disc and nearby supernovae in the Sun’s birth cluster. The ratio of $^{26}$ Al to its stable sibling $^{27}$ Al is nar- rowly peaked at a value of approximately $5 \times 10^{-5}$ in the CAIs found in the CV chondrites (Amelin et al., 2002; Connelly et al., 2012). Age determination of chondrules based on radio- genic $^{26}$ Mg yields chondrule ages that are systematically several million years younger than the CAIs (Villeneuve et al., 2009). However, direct measurement of ages of individual chondrules, based on the long-lived Pb-Pb system, indicates that the amount of $^{26}$ Al was originally lower in the chondrule-forming region compared to the region where CAIs condensed out (Connelly
| |
170805157/3
|
aration as chondrule-sized particles concentrate at km scales in strong disc turbulence. They found that the gas ionisation by de- cay of $^{26}$ Al must be reduced by several orders of magnitude in order to build up the break-down electric field.
The elevated densities needed for chondrule formation make an intriguing link to planetesimal formation by particle concen- tration by the streaming instability and gravitational collapse (Johansenetal.,2014). In this paper we therefore explore light- ning discharge and chondrule formation in dense pebble environ- ments. We show that secondary electrons released as positrons from the decay of $^{26}$ Al hammer through dust surfaces leads to positive charging of the pebbles. Charge separation by radial drift is nevertheless very slow inside dense pebble filaments and we find that the pebble current cannot compete with the neu- tralising current, even if abundance of $^{26}$ Al, and hence the gas conductivity, is decreased by an order of magnitude. Therefore we consider instead charge separation driven by the current of positrons released by the decay of $^{26}$ Al. Positrons stream out of dense pebble structures and cause a large-scale negative charg- ing of both circumplanetesimal discs and the sedimented mid- plane layer of pebbles in the protoplanetary disc. This charging is related to a direct-charging nuclear battery where a radioac- tive source charges a capacitor. The electric field can build up values where the electrons ionise the hydrogen molecule and cause lightning discharge. We construct simple models of the circumplanetesimal pebble disc and the protoplanetary disc en- vironments. We show that the e ffi ciency of converting the en- ergy of the decay of $^{26}$ Al to photons released in lightning dis- charge can be high in both these settings, although many uncer- tainties remain in the modelling of circumplanetesimal discs and the charging of filamentary pebble structures with density varia- tion on many scales.
The paper is organised as follows. In Section 2 we review the physics of lightning discharge and show by simple analytical ar- guments that charge separation driven by di ff erential drift of the pebbles relative to the dust and gas is orders of magnitude slower than the neutralising current. This conclusion is supported by thorough calculations of the equilibrium charge on gas, dust and pebbles presented in Appendices A–C, where we consider peb- ble charging by emission and absorption of positrons from the decay of $^{26}$ Al and release of secondary electrons after positron absorption (this mechanism is elaborated on in Appendix D). In Section 3 we analyse the positron current emanating from pebble filaments and sheets. We show that the positron current is higher than the neutralising current when the dust-to-gas ratio is above approximately 50. We discuss in Section 4 the operation of the nuclear battery e ff ect in circumplanetesimal discs and in the sed- imented mid-plane layer of pebbles. In the following Section 5 we demonstrate that the cooling rates of chondrules formed in dense pebble regions agrees with the long cooling times inferred from crystal growth experiments. Finally we summarise our re- sults in Section 6.
## 2. The problem with lightning discharge
In this section we review the physics of lightning discharge and demonstrate how charge separation by pebble sedimentation or di ff erential radial drift between pebbles charged oppositely of dust grains and gas drives currents in protoplanetary discs that are many orders of magnitudes lower than the neutralising gas current.
## 2.1. Lightning discharge
Lightning discharge requires the build up of an electric field $E$ strong enough that the accelerated electrons ionise the hydrogen molecule. This leads to a cascade e ff ect where the additional electrons increase the conductivity and hence the charge sepa- ration is neutralised in thin discharge channels. The breakdown electric field is given in Desch&Cuzzi(2000) as
$$
E_{\rm BD} = 85 f_{\rm BD} \mathopen{}\mathclose\bgroup\originalleft(\frac{\rho_{\rm g}}{10^{-7}\,{\rm
kg\,m^{-3}}} \aftergroup\egroup\originalright) \mathopen{}\mathclose\bgroup\originalleft(\frac{e_{\rm ion}}{60\,{\rm eV}} \aftergroup\egroup\originalright) \,{\rm
V\,m^{-1}} \, .
$$
Here we have normalised to a nominal gas density value $\rho_{\rm g}$ for the asteroid belt region in the primordial state of the solar protoplanetary disc and to an ionisation potential of $e_{\rm ion}=60$ eV. The ionisation energy of H $_2$ is 15.6 eV, but Desch&Cuzzi (2000) argued that impact energies below 60 eV lead to disso- ciation of the hydrogen molecule. Okuzumi & Inutsuka (2015) used a distribution function of electron energies and found that the breakdown of the hydrogen molecule occurs already at elec- tron energies around 5 – 10 eV; hence we multiplied equation (1) by a factor $f_{\rm BD}$ that we will assume in this paper to be 0.1.
The breakdown electric field can only be built up if the neutralising current, driven by growing electric field, remains smaller than the current driven by the motion of the charged solid particles. The gas has an intrinsic conductivity that is a function of the density and mobility of ions and electrons. Ionisation in the dense mid-plane layer is dominated by the decay of short- lived radionuclides, most importantly by release of positrons in the decay of $^{26}$ Al (Desch & Cuzzi, 2000). External sources of ionisation, such as cosmic rays, can be ignored in the dense mid- plane in the asteroid formation region.
## 2.2. Ion and electron densities
The gas conductivity depends strongly on the number densities of ions and electrons, which are in turn dictated by adsorption onto dust grains. We consider throughout this paper for simplic- ity a two-component model for the solid particles. Component 1 consists of microscopic dust grains that are the main absorbers of electrons and ions, while component 2 consists of macro- scopic (dust aggregate) pebbles that can form chondrules after flash heating. The number densities of electrons, ions, positrons, small grains and pebbles are $n_{\rm e}$, $n_{\rm i}$, $n_\beta$, $n_1$ and $n_2$. The two species of solid particles have masses $m_1$ and $m_2$, radii $R_1$ and $R_2$, and charges (in units of the elementary charge $e$) $q_1$ and $q_2$. The gas is characterised by a neutral density $n_{\rm n}=\rho_{\rm g}/m_{\rm n}$, where $m_{\rm n}$ is the neutral (hydrogen molecule) mass. Cross sections are denoted $\sigma$, with single subscript used to indicate the target species or dou- ble subscript to indicate that both collision partners matter for the cross section.
Positrons released by the decay of $^{26}$ Al lose their energy to ionisation of the hydrogen molecule (with ionisation poten- tial $E_{\rm ion}$) and to stopping in solid particles. The positron kinetic energy has an average value of $T_\beta=0.66$ MeV and we take the energy loss $E_{\rm ion}=37$ eV per ionisation (Glassgold, 1995), yielding in the absence of dust around 18,000 ionisations be- fore the positron finally annihilates with a bound electron to cre- ate a $\gamma$ -ray. This $\gamma$ -ray, in turn, has a very long stopping length and its e ff ects on the mid-plane ionisation can be ignored. A more typical end to the positron trajectory is penetration in a solid particle. The stopping column density of positrons in sil- icon is $\varSigma_\beta \approx
3\,{\rm kg\,m^{-2}}$and the stopping length is$\chi_{\rm s} \approx 1.3\,{\rm
mm}$(see Section 3.1). We use the terminology$R_1^*$,$R_2^*$,$m_1^*$and$m_2^*$ to
| |
170805157/9
|
could operate to thermally process solids: in pebble discs orbit- ing young planetesimals and in the sedimented mid-plane layer of pebbles in the main protoplanetary disc.
## 4.1. Nuclear battery in a circumplanetesimal pebble disc
The high abundance of sodium within chondrules can be under- stood if chondrule precursors were heated in an environment that was several orders of magnitude denser than the Roche density at the distance of the asteroid belt from the Sun (Alexanderetal., 2008). Such dense regions should be undergoing gravitational collapse. The gravitational collapse phase is short, on the or- der of a few orbital periods (Nesvorn Jansson & Johansen, 2014). How can chondrules in chondrites reflect formation conditions that were prevalent for only the last tiny fraction of the free life of a pebble in a protoplanetary disc before it was incorporated in a planetesimal? Actually, densi- ties higher the Roche limit do not necessarily imply gravita- tional collapse, if the collapsing region is rotationally supported. Johansen&Lacerda(2010) demonstrated how pebbles accreted onto protoplanets enter a prograde particle disc in stable orbit around the central protoplanet. The Roche density in such a par- ticle disc will be orders of magnitude higher than the Roche den- sity that results from the tidal force of the central star alone. Could the high particle densities required for chondrule heating have been sustained for hundreds of thousands of years in thin pebble discs orbiting around planetesimals and protoplanets?
A circumplanetesimal particle disc will not evolve in isola- tion from its surroundings. The sub-Keplerian gas acts as a fric- tion on the outermost pebbles, unbinding them from the grav- ity of the planetesimal so that they mix with other pebbles in the protoplanetary disc and participate in new generations of planetesimal formation and pebble accretion. Planetesimals that enter eccentric orbits, e.g. perturbed by the growing protoplan- ets (Johansen et al., 2015), experience even stronger gas head- winds that could erode o ff pebbles even more e ffi ciently. A large fraction of the chondrules that are flash-heated in a cir- cumplanetesimal disc are nevertheless expected to become ac- creted onto the central planetesimal. Radiometric dating of indi- vidual chondrules, by the long-lived Pb-Pb chronometer, shows that chondrules in the same chondrite display a wide range of ages (Connellyetal.,2012). This age di ff erence could reflect ei- ther that the parent bodies of those meteorites accumulated late from a mixture of chondrules that formed in other circumplan- etesimal discs or that the chondrules that are accreted through a circumplanetesimal disc experience e ffi cient mixing between chondrules formed at various epochs.
We can estimate the column density of a circumplanetesimal disc from the assumption that the disc evolves towards a state that is marginally unstable to self-gravity. In this view the cir- cumplanetesimal disc accretes mass onto the central planetes- imal by large-scale gravitational stresses when onfalling peb- bles lower the Safronov-Toomre $Q$ parameter to below unity (Safronov, 1960; Toomre, 1964). The marginal case of $Q=1$ gives a column density of
$$
\varSigma_{\rm p} = \frac{c \varOmega_{\rm pla}}{\pi G} \, .
$$
Here $c$ is the particle velocity dispersion and $\varOmega_{\rm pla}(r)$ is the Keplerian frequency profile of the circumplanetesimal disc. The velocity dispersion $c$ is generally not known. Completely elas- tic particle collisions cause an exponential transfer of the rel- ative shearing motion of the particles to random particle mo- tion. The evolution of the random motion stops when the coe ffi - cient of restitution drops at higher collision speeds (Goldreich& Tremaine,1978). Estimates of the velocity dispersion in Saturn’s rings motivates setting this transition in the elasticity of the col- lisions at $0.001\,{\rm m\,s^{-1}}$. However, the flu ff y silicate pebbles that form in the inner regions of protoplanetary discs may have very di ff erent energy dissipation in inelastic collisions than the ice particles of Saturn’s rings and hence we keep $c$ as a free param- eter in the circumplanetesimal disc model. The profiles of parti- cle column density, particle scale-height and mid-plane particle density, all calculated for $Q=1$, come out as
$$
\begin{aligned}
\varSigma_{\rm p} &=& 3.2\,{\rm kg\,m^{-2}}\, \mathopen{}\mathclose\bgroup\originalleft(\frac{r}{100\,R_{\rm
pla}} \aftergroup\egroup\originalright)^{-3/2} \mathopen{}\mathclose\bgroup\originalleft(\frac{c}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright) \mathopen{}\mathclose\bgroup\originalleft(
\frac{R_{\rm pla}}{10^2\,{\rm km}} \aftergroup\egroup\originalright)^{3/2} \, , \\
H_{\rm p} &=&
130\,{\rm km}\,\mathopen{}\mathclose\bgroup\originalleft(\frac{r}{100\,R_{\rm pla}} \aftergroup\egroup\originalright)^{3/2} \mathopen{}\mathclose\bgroup\originalleft(
\frac{c}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright) \, , \\
\rho_{\rm p} &=& 1.1 \times 10^{-3}\,{\rm kg\,m^{-3}} \mathopen{}\mathclose\bgroup\originalleft(
\frac{r}{100\,R_{\rm pla}} \aftergroup\egroup\originalright)^{-3}
\mathopen{}\mathclose\bgroup\originalleft(\frac{R_{\rm pla}}{10^3\,{\rm km}} \aftergroup\egroup\originalright)^{3/2} \, .
\end{aligned}
$$
$$
\begin{aligned}
\varSigma_{\rm p} &=& 3.2\,{\rm kg\,m^{-2}}\, \mathopen{}\mathclose\bgroup\originalleft(\frac{r}{100\,R_{\rm
pla}} \aftergroup\egroup\originalright)^{-3/2} \mathopen{}\mathclose\bgroup\originalleft(\frac{c}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright) \mathopen{}\mathclose\bgroup\originalleft(
\frac{R_{\rm pla}}{10^2\,{\rm km}} \aftergroup\egroup\originalright)^{3/2} \, , \\
H_{\rm p} &=&
130\,{\rm km}\,\mathopen{}\mathclose\bgroup\originalleft(\frac{r}{100\,R_{\rm pla}} \aftergroup\egroup\originalright)^{3/2} \mathopen{}\mathclose\bgroup\originalleft(
\frac{c}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright) \, , \\
\rho_{\rm p} &=& 1.1 \times 10^{-3}\,{\rm kg\,m^{-3}} \mathopen{}\mathclose\bgroup\originalleft(
\frac{r}{100\,R_{\rm pla}} \aftergroup\egroup\originalright)^{-3}
\mathopen{}\mathclose\bgroup\originalleft(\frac{R_{\rm pla}}{10^3\,{\rm km}} \aftergroup\egroup\originalright)^{3/2} \, .
\end{aligned}
$$
$$
\begin{aligned}
\varSigma_{\rm p} &=& 3.2\,{\rm kg\,m^{-2}}\, \mathopen{}\mathclose\bgroup\originalleft(\frac{r}{100\,R_{\rm
pla}} \aftergroup\egroup\originalright)^{-3/2} \mathopen{}\mathclose\bgroup\originalleft(\frac{c}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright) \mathopen{}\mathclose\bgroup\originalleft(
\frac{R_{\rm pla}}{10^2\,{\rm km}} \aftergroup\egroup\originalright)^{3/2} \, , \\
H_{\rm p} &=&
130\,{\rm km}\,\mathopen{}\mathclose\bgroup\originalleft(\frac{r}{100\,R_{\rm pla}} \aftergroup\egroup\originalright)^{3/2} \mathopen{}\mathclose\bgroup\originalleft(
\frac{c}{0.1\,{\rm m\,s^{-1}}} \aftergroup\egroup\originalright) \, , \\
\rho_{\rm p} &=& 1.1 \times 10^{-3}\,{\rm kg\,m^{-3}} \mathopen{}\mathclose\bgroup\originalleft(
\frac{r}{100\,R_{\rm pla}} \aftergroup\egroup\originalright)^{-3}
\mathopen{}\mathclose\bgroup\originalleft(\frac{R_{\rm pla}}{10^3\,{\rm km}} \aftergroup\egroup\originalright)^{3/2} \, .
\end{aligned}
$$
In these expressions we have normalised the radial distance from the planetesimal, $r$, by 100 times the planetesimal radius, $R_{\rm pla}$. We plot the radial profiles of pebble column density, ra- tio of disc mass to planetesimal mass, pebble scale-height and pebble mid-plane density for a planetesimal of radius 100 km and a protoplanet of radius 1000 km in Figure 4. The conditions for the nuclear battery depend strongly on the adopted value of the random pebble speed $c$ as well as on the planetesimal size. Discs with random pebble speed in the approximate range be- tween 0.01 m / s to 0.1 m / s have column densities close to the stopping column of positrons between 100 and 1000 planetesi- mals radii where most of the disc mass resides.
The ratio of the circumplanetesimal disc mass to the mass of the central planetesimal determines the maximum production e ffi ciency of chondrules. The top right panel of Figure 4 shows that the protoplanet of radius 1000 km has only a small mass fraction in the circumplanetesimal disc and hence can only pro- duce a relatively small amount of chondrules relative to its mass. This may nevertheless not necessarily be in conflict with the high mass fraction of chondrules in chondrites, if these chondrules were produced in pebble discs around larger protoplanets. As the outer edge of the disc expands by angular momentum ex- change with the inner disc, the chondrules formed in the circum- planetesimal disc are recycled into the protoplanetary disc and could contribute significantly to the mass reservoir from which the smaller chondrite parent bodies formed, even if the produc- tion e ffi ciency relative to the protoplanet mass is low.
Importantly, the bottom right panel of Figure 4 illustrates how the pebble density inside of the Hill radius of the planetes- imal reaches several orders of magnitude above the Roche den- sity in the protoplanetary disc. Thus the mid-plane conditions around both the planetesimal and the protoplanet are right in the middle of the constraints inferred by Alexander et al. (2008) and Alexander & Ebel (2012). The circumplanetesimal discs that formed in the simulations by Johansen & Lacerda (2010) had sizes around 10% of the Hill radius, or 100 times the plan- etesimal radius. These discs would nevertheless not form around small planetesimals, due to the headwind of the gas. Therefore it is doubtful whether small planetesimals can host extensive cir- cumplanetesimal discs, unless the gravitational collapse in itself leads to the formation of a compact, rotationally supported peb- ble clump that subsequently expands by viscous stresses into its Hill radius. Such a dense disc expanding outwards could have enough density to withstand the friction from the sub-Keplerian gas. This model is nevertheless very speculative and this illus-
| |
170805157/8
|
To get a better handle on realistic values of the positron current and the charging e ffi ciency we perform a Monte Carlo simula- tion of positron emission and absorption in a dense particle fila- ment. We use a Gaussian profile for the density of particles, with peak mass loading $\epsilon_2$ relative to the gas, embedded in a constant background particle density with mass loading $\epsilon_1$,
$$
\rho_{\rm p}=\epsilon_1 \rho_{\rm g}+\epsilon_2 \rho_{\rm g}
\exp[-(x^2+z^2)/(2 W^2)] \, .
$$
Note that $\epsilon_1$ and $\epsilon_2$ do not represent respectively small and large particles (and hence are unrelated to $Z_1$ and $Z_2$ employed in Section 2); from the view of positron emission the particle size is insignificant, as long as the particles are smaller than approxi- mately 1 mm, so that the positron can freely escape from the in- terior. We create positrons at random positions that are weighted with the ratio of the local particle density $\epsilon_1+\epsilon_2$ to the total dust mass. The positrons are given random 3-D directions and are ab- sorbed after interacting with the characteristic stopping column of $\varSigma_\beta$ in gas and particles. In Figure 1 we present the results of a simulation with $\rho_{\rm
g}=10^{-7}\,{\rm kg\,m^{-3}}$,$\epsilon_1=0.1$,$\epsilon_2=1000$and$W=20$km. The center of the filament charges rapidly negative, at 100 electron charges per cubic meter per second. The mea- sured current density streaming out of the filament is shown in Figure 2, as a function of the solids-to-gas ratio (a measure of the distance from the centre of the filament). The current den- sity is above$3
\times 10^{-14}$C m$^{-2}$s$^{-1}$in most of the region and peaks at a value of$4.5 \times 10^{-14}$C m$^{-2}$s$^{-1}$, near the characteristic value found in equation (30). The rate of change of the electric field is related to the current density through the simple expres- sion$\dot{E} =
\epsilon_0^{-1} J$, with vacuum permittivity$\epsilon_0$. The right axis of Figure 2 shows that the breakdown value of the electric field is reached within a characteristic time-scale of only$10^3$ s. The fil- ament thus quickly builds up an electric field strength near the breakdown value and the continued charging by positron release is neutralised by lightning, keeping the electric field around the breakdown value.
The e ffi ciency of the nuclear battery e ff ect can be calculated from the net charging. We perform here a series of experiments where we cover a pebble structure and its surroundings by a fixed grid and monitor the emission and absorption of positrons in each cell. We consider a gas density of $\rho_{\rm g}
= 10^{-7}\,{\rm kg\,m^{-3}}$and a box size of$10^5$km to capture the positron stopping length in the gas of approximately$3 \times 10^4$km. The dust-to-gas ratio was cho- sen to be$\epsilon_1=0.1$and the pebble-to-gas ratio varied from$\epsilon_2=1$to$\epsilon_2=10^3$. The pebble structure scale is fixed at$3
\times 10^3$km, larger than the example shown in in Figure 1 because we must be careful to capture here the large length-scales of positron ab- sorption in the gas. The transition between being thin and thick to positron emission happens then at a pebble mass loading of$\epsilon_2 \approx 10$.
Releasing a total of $N_\beta$ positrons inside the pebble structure and measuring a net charge of $Q_j$ on the grid, where $j$ is the grid index, yields an e ffi ciency of
$$
\eta_{\rm c} = \frac{\sum_j |Q_j|}{N_\beta} \, .
$$
Here we sum only over regions inside of the pebble structure, defined as $\rho_2
> \epsilon_1 \rho_{\rm g}$. This exclusion is necessary since positrons released near the boundary of the computational domain are not replaced with a counterstreaming flux and hence appear to have high charging e ffi ciency. A uniform particle density would yield$\eta_{\rm c} \approx 0$, since all cells would have equally many emitted and absorbed positrons. The e ffi ciency of the nuclear battery e ff ect is shown in Figure 3 as a function the pebble mass loading$\epsilon_2$. We
Fig.3. Charging e ffi ciency (solid lines) and current density (dashed lines) as a function of the mass loading in pebbles, for both 1-D fila- ments (blue lines) and 2-D sheets (red lines), for a characteristic struc- ture width of 3,000 km. The charging e ffi ciency is high at low pebble densities when the positrons can escape freely from the structure, but falls for high pebble densities due to self-absorption of positrons within the structure. The current behaves oppositely: at low pebble densities the current is low, since the structure is thin to its own positron emis- sion, while the current density plateaus around the characteristic value of $3.6 \times 10^{-14}\,{\rm
C\,m^{-2}\,s^{-1}}$ for high pebble densities.
consider both 1-D filaments and 2-D sheets. The most e ffi cient charging is obtained when the pebble density is low, as most positrons can escape such dilute pebble structures. The current is only $10^{-14}$ C m $^{-2}$ s $^{-1}$ for low values of the pebble density, but plateaus around the characteristic value of $3.6 \times 10^{-14}\,{\rm
C\,m^{-2}\,s^{-1}}$when the pebble structure transitions from being thin to thick to its own positron emission (at$\epsilon_2 \approx 10$).
## 4. Application to protoplanetary discs
The nuclear battery e ff ect identified and described in Section 3 provides a pathway to convert the decay energy of $^{26}$ Al to light- ning discharge that can flash heat solids. The total energy present in positrons released by $^{26}$ Al of decay rate $r_\beta=5\times10^5\,{\rm kg^{-1}\,s^{-1}}$, energy $T_\beta = 0.66$ MeV and decay constant $\tau_{26} = 1.03$ Myr is
$$
E_{26} = r_\beta T_\beta \tau_{26} \approx 1.7 \times 10^6\,{\rm J\,kg^{-1}}
\, .
$$
The heating and partial melting of chondrule precursors re- quires approximately $1.6 \times 10^6\,{\rm J\,kg^{-1}}$. The positrons thus release, over the life-time of $^{26}$ Al of a few million years, approximately the energy needed to melt all rock in the protoplanetary disc 2. Transferring the positron energy to the electric field and then to lightning discharge can therefore lead to flash heating of a significant fraction of the solids in the disc, if the energy transfer and heating e ffi ciencies are both high. We discuss here two settings where the nuclear battery e ff ect
2 Note that the full heating capacity of $^{26}$ Al trapped inside a planetes- imal is much higher, at 4 MeV per decay compared to 0.66 MeV that goes into the kinetic energy of the positron, due to eventual annihilation of the positron to produce a $\gamma$ -ray.
| |
190408596/5
|
FIG. 4: (a) Estimated depairing current using the fast relax model as a function of the operating temperature for both NbN and WSi devices and (b) the switching to depairing current ratio (constriction factor) for all the tested devices as a function of the fraction of superconductor transition temperature.
nanowires with respect to biasing at different temper- ature conditions. This measurement was done to com- pare the temperature dependence of the depairing cur- rent with the theoretical predictions. The NbN nanowires resonance frequencies were collected starting from the base temperature 1.05 K up to 7.00 K, which is more than 80% of the superconductor transition temperature, measured to be 8.65 K, while the WSi devices were mea- sured up to 2.45 K, which corresponds to 70% of their $T_c$ of 3.50 K. The constriction factor drops with increasing temperature for both NbN and WSi devices. This ef- fect might be due to local defects in the nanowire struc- ture: if the weakest constriction in the nanowire has a lower transition temperature then its local switching cur- rent would drop faster than the depairing current for the whole nanowire, with increasing operating temperature. This observation deserves future investigation, since it could shed light on the possibility of SNSPD operation at elevated temperatures.
In total, we tested one die with two NbN device ge- ometries (widths of 120 and 140 nm) and two identical dies with five WSi device geometries each (widths of 55, 80, 120, 160 and 200 nm). The measured switching cur- rents and the estimated critical depairing currents based on the fast and the slow relaxation models are collected in Table I. In Fig. 4 a we report the trend of the devices’ critical depairing currents with respect to different tem- peratures. We estimate the zero temperature depairing current $I_{\text{dep}}(0)$ by fitting the measured temperature de- pendence of $I_{\text{dep}}(T)$ to the function defining the temper- ature dependence of the numerical solution to the Us- adel equations. These estimated values for $I_{\text{dep}}(0)$ are collected in Table I. For comparison, we also calculated the theoretical critical depairing current at zero temper- ature according to Kupryianov and Lukichev model [22], denoted as:
$$
I_{\text{dep}}^{\text{KL}}(0) = 1.491~e~N(0)~[\Delta(0)]^{3/2}\sqrt{D/\hbar}~w d
$$
where $e$ is the electron charge, $N(0) = (2e^2DR_\text{sq}d)^{-1}$ is the single-spin electron density of states at Fermi level in the normal state, $\Delta(0) = 1.764~k_BT_c$ is the super- conducting gap at zero temperature, $D$ is the diffusion coefficient, $R_\text{sq}$ is the square resistance, and $w$ and $d$ are width and thickness of the nanowire, respectively. In or- der to calculate these values, we measured the diffusion coefficient for WSi, while for NbN, we used a value found in literature.
In order to calculate these values, we measured the temperature dependence of the upper critical magnetic field ($B_{c2}$) and extracted information on material prop- erties of the WSi thin film. The electron diffusion coef- ficient $D$ was obtained from the slope of the $B_{c2}$ vs $T$ curve. In the limit of a dirty superconductor, the elec- tron diffusivity $D$ can be expressed as follows, based on [32],
$$
D = \frac{1.097}{\Big[-\frac{dB_{c2}(T)}{dT}\Big]_{T=T_c}},
$$
where the diffusion coefficient $D$ has dimensions of $[\text{cm}^2\text{s}^{-1}]$, the upper critical magnetic field $B_{c2}$ has di- mensions of [T] and the temperature has dimensions of [K].
The external magnetic field was applied perpendic- ular to the surface of the film and $B_{c2}$ was defined as the field at which the resistance of the film becomes half of the normal state value. The calculated value for the electron diffusion coefficient, based on equation (6), is 0.74 $\text{cm}^2/\text{s}$ for the 7 nm thick WSi film. The Ginzburg- Landau coherence length, $\xi_{GL}(0)$, at $T=0$ can be ex-
| |
190408596/2
|
FIG. 1: Scanning electron micrographs of NbN CPW resonator used in experiment. (a) The narrow, meandered nanowire CPW is placed between two wide, 50 $\Omega$ leads (also in CPW configuration), forming a transmission-line resonator. (b) Transition from the 50 $\Omega$ lead to the k $\Omega$ nanowire. (c) Zoomed-in view of the nanowire CPW.
bias-tee and amplifier.
We present the dependence of the measured depair- ing current on the width of the nanowire resonators as well as the operating temperature. An important obser- vation is that $C$ reduces for higher operating tempera- tures for both polycrystalline NbN and amorphous WSi devices, which has significant consequences for design of high-performance SNSPDs at elevated temperatures. There is also an indication that narrower nanowires achieve a lower $C$, which may point to nanowire edge roughness due to fabrication imperfections.
## DEVICE DESIGN AND FABRICATION
The nanowire resonators are designed in a CPW [26, 27] to avoid electromagnetic coupling within the me- ander and to allow for simplified impedance engineer- ing. The resonance is set up by means of the impedance mismatch between the transmission line and the narrow nanowire. This approach simplifies current biasing of the nanowire. The devices were designed in order to have the resonant frequency at roughly 2 GHz, so that the microwave period ($\tau_{\text{exp}} \approx$ 500 ps) is much larger than the relaxation time of the superconducting order param- eter $\tau_{\text{s}}$ for both WSi [28] and NbN [29]. An estimate of the relaxation time of the order parameter is given by
FIG. 2: Schematics for the setup for both the (a) reflection and (b) transmission type measurements. The THRU devices were 50 ohm superconducting CPW fabricated on the same chip used for calibration purposes.
$\tau_{\text{s}} = \hbar / k_B (T_c-T)$,
so for NbN films ($T_c=$ 8.65 K) the or- der parameter relaxation time is 1 ps, while for WSi films ($T_c=$ 3.50 K) it is 3.1 ps, at a temperature of 1.05 K. At the highest temperature investigated ($0.8T_c$ for NbN and $0.7T_c$ for WSi) the order parameter relaxation time is 4.6 ps and 7.3 ps, for NbN and WSi films, respectively.
The devices were fabricated from a 6 nm thick NbN film and from a 7 nm thick WSi film. NbN film was sputter deposited on a 4-inch silicon wafer with a 300 nm thick thermal oxide layer [25]. WSi was sputter deposited on 4-inch silicon wafer with a 240 nm thick thermal ox- ide layer and was passivated with a 15 nm thick silicon dioxide (SiO $_2$) film. All the devices and pad structures were patterned using $125\,\mathrm{kV}$ electron beam lithography with gL2000 positive tone resist [26]. The patterns were then transferred into NbN and WSi by CF $_4$ reactive ion etching. A layer of HSQ was spun on the dies after fab- rication, for passivation.
| |
190408596/6
|
\begin{tabularx}{\textwidth}{ccYYYYYYYYcYY}
\hline
\multicolumn{2}{>{\hsize=\dimexpr2\hsize+2\tabcolsep+\arrayrulewidth\relax}Y}{\textbf{Device}} & \multicolumn{2}{>{\hsize=\dimexpr2\hsize+2\tabcolsep+\arrayrulewidth\relax}Y}{\textbf{Fast Relax Approximation}} & \multicolumn{2}{>{\hsize=\dimexpr2\hsize+2\tabcolsep+\arrayrulewidth\relax}Y}{\textbf{Fast Relax Numerical}} & \multicolumn{2}{>{\hsize=\dimexpr2\hsize+2\tabcolsep+\arrayrulewidth\relax}Y}{\textbf{Slow Relax Approximation}} & \multicolumn{3}{>{\hsize=\dimexpr3.5\hsize+2\tabcolsep+\arrayrulewidth\relax}Y}{\textbf{Measured}} & \multicolumn{2}{>{\hsize=\dimexpr2\hsize+2\tabcolsep+\arrayrulewidth\relax}Y}{\textbf{Estimated}} \\
Material & Width & $I_{\text{dep}}$ & Fit $R^2$ & $I_{\text{dep}}$ & Fit $R^2$ & $I_{\text{dep}}$ & Fit $R^2$ & $I_{\text{sw}}$ & $C$ & $\Big\{\frac{\mathcal{L}_{\text{k}}(q,T)}{\mathcal{L}_{\text{k,0}}(T)} \Big\}_{\text{sw}}$ & $I_{\text{dep}}(0)$ & $I_{\text{dep}}^{\text{KL}}(0)$\\[.5\normalbaselineskip] \hline
WSi & 55 nm & 4.40 & 0.9942 & 4.67 & 0.9990 & 3.32 & 0.7413 & 2.25 & 0.54 & 1.107 & 5.31* & 6.47 \\
\rowcolor{gray!10}WSi & 55 nm & 4.29 & 0.9924 & 4.59 & 0.9984 & 3.09 & 0.7998 & 2.13 & 0.49 & 1.085 & 5.09 & 6.47 \\
WSi & 80 nm & 7.58 & 0.9808 & 8.30 & 0.9962 & 4.74 & 0.9636 & 3.25 & 0.43 & 1.055 & 9.72 & 10.68 \\
\rowcolor{gray!10}WSi & 80 nm & 9.22 & 0.9955 & 9.81 & 0.9995 & 6.05 & 0.9838 & 4.75 & 0.52 & 1.094 & 11.66 & 10.68 \\
WSi & 120 nm & 14.62 & 0.9930 & 15.60 & 0.9996 & 9.47 & 0.9801 & 7.25 & 0.50 & 1.090 & 18.72 & 17.41 \\
\rowcolor{gray!10}WSi & 120 nm & 14.82 & 0.9940 & 16.25 & 0.9961 & 9.55 & 0.9834 & 6.75 & 0.46 & 1.066 & 20.31 & 17.41 \\
WSi & 160 nm & 20.76 & 0.9980 & 21.71 & 0.9986 & 13.82 & 0.9856 & 12.25 & 0.59 & 1.152 & 26.07 & 23.46 \\
\rowcolor{gray!10}WSi & 160 nm & 21.16 & 0.9970 & 21.98 & 0.9984 & 14.44 & 0.9750 & 13.50 & 0.64 & 1.179 & 25.98 & 23.46 \\
WSi & 200 nm & 27.65 & 0.9954 & 28.06 & 0.9993 & 21.00 & 0.7813 & 20.50 & 0.74 & 1.313 & 33.26 & 30.28 \\
\rowcolor{gray!10}NbN & 120 nm & 38.19 & 0.9975 & 38.78 & 1.0000 & 27.05 & 0.9239 & 26.50 & 0.69 & 1.280 & 42.07 & 43.30\\
NbN & 140 nm & 46.93 & 0.9970 & 47.67 & 0.9999 & 33.09 & 0.9295 & 32.50 & 0.69 & 1.280 & 51.47 & 50.52\\ \hline
\end{tabularx}
TABLE I: Table representing the results obtained at base temperature (1.05 K). (*) The estimated depairing current at zero Kelvin for the 55 nm wide nanowire is estimated fitting only two points. Switching currents ($I_{\text{sw}}$) were extracted from IV curves, measured at a rate of several minutes per sweep.
tracted from the following equation:
$$
B_{c2}(T) = \frac{\Phi_0}{2~\pi~\xi(T)^2},
$$
where $\Phi_0 = h/2e$ is the magnetic-flux quantum and $e$ is the electron charge.
In the limit of a dirty superconductor, a linear ex- trapolation of the measured $B_{c2}(T)$ down to $T = 0$, over- estimates the real upper critical field at zero tempera- ture and consequently underestimates the superconduct- ing coherence length. A more realistic value of $B_{c2}(0)$ is given by
$$
B_{c2}(0) = 0.69~T_c~\Big[-\frac{dB_{c2}(T)}{dT}\Big]_{T=T_c}.
$$
Using this value of $B_{c2}(0)$ in equation (7) the cal- culated coherence length is 9.62 nm for the 7 nm thick WSi film. For the 6 nm thick NbN film, we considered a diffusion coefficient of $D = 0.5~\text{cm}^2/\text{s}$ as used by Zhao et al. [26] and estimated a coherence length at zero tem- perature $\xi_{GL}(0)$ of 5.01 nm.
We measured the constriction factor, $C(T) = I_{\text{sw}}(T) / I_{\text{dep}}(T)$, which is the ratio between the switch- ing and depairing current, at different temperature con- ditions for all the devices tested. This ratio can be con- sidered as the quality of the nanowire itself. Shown in Fig. 4 b, the ratio of currents suggests a decrease of qual- ity of the devices with increasing temperature.
For the WSi devices, since five geometries were stud- ied, we were able to show the dependence of the depairing current on the device width. As Fig. 5 shows a linear fit to the depairing current estimated at different temper- atures seems to suggest that the effective widths of the nanowires might be reduced from the measured widths
FIG. 5: Depairing current at different temperatures with respect to increasing resonator width for NbN and WSi devices. The linear fit for the WSi devices shows the presence of an offset.
(SEM after etching) by an offset of $\sim 23$ nm. That effect could be caused by the loss of superconductivity in the edges of the nanowire due to scattering of particles dur- ing etching, or due to oxidation of the nanowire caused by exposure to the environment. It is worth noting that the offset is close to two times the superconducting co- herence length $\xi_{GL}$ of the WSi devices, so it is possible that poisoning of the edges of the nanowire during the fabrication process might have suppressed the supercon- ducting active area by roughly one coherence length on each side. More work is needed to conclusively determine the the cause of this observation.
| |
190408596/4
|
FIG. 3: The measured and the fitting functions for the resonance (a) magnitude and (b) phase responses at zero and near the switching bias current for the 120 nm wide NbN device. Fitting of the kinetic inductance ratio of the nanowire (c) using the fast (in blue) and the slow (in red) relaxation approximation models for the 120 nm wide NbN device. The shaded area represents the model’s accuracy of 1% and 0.5% for the fast and slow relaxation approximation models, respectively, according to Clem and Kogan [23]. In black, the numerical simulation of the kinetic inductance change using the fast relaxation model. It can be shown by the enlarged window that the numerical simulation fits the sample points better than the fast relax approximation model. The estimated depairing current from the models is 38. 19 µ A for the fast relaxation approximation, 38. 78 µ A for the fast relaxation numerical simulation and 27. 05 µ A for the slow relaxation.
## RESULTS
We measured the resonant frequencies of nanowire devices with widths of typical SNSPDs (50-200 nm) using both NbN and WSi thin films.
The resonance features of an NbN, 120nm wide de- vice, measured at zero current and close to the switching bias current ($\text{I}_{\text{bias}} = \SI{26}{\micro\ampere}, \text{I}_{\text{sw}} = 27 \pm .5 ~\SI{}{\micro\ampere}$) are shown in Fig. 3. The fit of the models described in section IV is shown in red for both the transmission (Fig. 3 a, where we use (2) to fit the magnitude) and reflection (Fig. 3 b, where we use (1b) to fit the phase) measurements. For the phase analysis, we found it best to normalize the phase data with respect to the phase of the resonator while in non-superconducting state, i.e. biasing the de- vice above its switching current. The resonant peaks ob- tained using the two different methods of reflection and transmission match within 0.5%; however, from the good- ness of the two fits, we decided to prioritize the analysis of the phase in reflection method as it is, in general, less noisy and requires fewer free parameters to perform the fit. From this point onward we only refer to data col- lected from the phase response of the resonator in reflec- tion mode.
The kinetic inductance ratios, obtained by the mea- sured resonance frequencies as,
$$
y(\text{I}_{\text{bias}},T) = \frac{\mathcal{L}_{\text{k}}(q,T)}{\mathcal{L}_{\text{k,0}}(T)} = \Bigg[\frac{\omega_r(\text{I}_{\text{bias}} = 0,T)}{\omega_r(\text{I}_{\text{bias}},T)}\Bigg]^2 \text{,}
$$
are then plotted in Fig. 3 c a function of the bias current. The fast relaxation and slow relaxation models discussed by Clem and Kogan [23] have been fitted to the data, where the only free parameter is the depairing current $I_{\text{dep}}$ of the nanowire. The estimated depairing current for each model can be found in the caption of Fig. 3.
It is immediately clear from Fig. 3 c that the fast re- laxation model provides a better fit of the experimental data than the slow relaxation model. Moreover, the de- pairing current evaluated using the latter model appears to be unreliable since the model predicts depairing cur- rents just above the measured switching currents. Due to fabrication imperfections, the measured switching cur- rents in SNSPDs are typically significantly below the de- pairing current since the switching current is set by the weakest point along the nanowire, typically referred to as a constriction. The depairing current, however, is an average characteristic of the nanowire, hence a switching current approaching the depairing current would suggest a ”perfect” nanowire. By removing the highest bias cur- rent points, it is possible to simulate a more constricted nanowire, while the measured depairing current should remain unchanged. Carrying out this exercise, the slow relaxation model does not predict constant values while the fast relaxation model is robust and provides depairing current estimates which are more consistent with theoret- ical models. With this, we conclude that our experiment falls into the fast relaxation regime, which has not been confirmed previously [24].
We measured the resonant frequency of the
| |
190408596/1
|
# Determining the depairing current in superconducting nanowire single-photon detectors
S. Frasca, 1, ∗
B. Korzh, 1, †
M. Colangelo, 2
D. Zhu, 2
A. E. Lita, 3
J. P. Allmaras, 1,4
E. E. Wollman, 1
V. B. Verma, 3
A. E. Dane, 2
E. Ramirez, 1
A. D. Beyer, 1
S. W. Nam, 3
A. G. Kozorezov, 5
M. D. Shaw, 1 and
K. K. Berggren 2
We estimate the depairing current of superconducting nanowire single-photon detectors (SNSPDs) by studying the dependence of the nanowires’ kinetic inductance on their bias cur- rent. The kinetic inductance is determined by measuring the resonance frequency of resonator- style nanowire coplanar waveguides both in transmission and reflection configurations. Bias current dependent shifts in the measured resonant frequency correspond to the change in the kinetic induc- tance, which can be compared with theoretical predictions. We demonstrate that the fast relaxation model described in the literature accurately matches our experimental data and provides a valu- able tool for direct determination of the depairing current. Accurate and direct measurement of the depairing current is critical for nanowire quality analysis, as well as modeling efforts aimed at understanding the detection mechanism in SNSPDs.
## INTRODUCTION
Superconducting nanowire single-photon detectors (SNSPDs) [1] are established as a key technology for many applications, such as deep-space optical communi- cation, laser ranging and quantum science. This is due to their high efficiency ($>$ 90%) [2], wide wavelength sensi- tivity (from X-rays to mid-infrared) [3, 4], low dark count rate ($<$ 1 Hz) [5] and ultra-high timing resolution ($<$ 3 ps) [6].
In the last decade, widespread effort by the SNSPD community has improved the theoretical understanding of the detection mechanism in SNSPDs. Guided by ex- perimental measurements [7 – 14] and theoretical model- ing [15 – 20], it is currently understood that most fea- tures of photodetection in SNSPDs can be explained by a combination of Fano fluctuations [20] and vortex-based breaking of superconductivity [16]. More recently, the measurement of record low timing jitter [6] has led to a new effort in understanding the latency of SNSPDs [21] in order to predict the intrinsic timing jitter of these detec- tors. It is known that a precise estimate of the depairing current of a device is needed in order to match exper- imental results using these models. The most common way of estimating the depairing current is through the Kupryianov-Lukichev formula [22], which requires sev- eral independent material parameters such as the dif- fusion coefficient, sheet resistance, critical temperature and nanowire geometry. In this work, we demonstrate a direct method of accessing the depairing current by measuring the kinetic inductance change as a function of the bias current. This method relies on fitting the the- oretical dependence calculated by Clem and Kogan [23] where the depairing current is the single free fitting pa- rameter. Having access to a direct measurement of the depairing current enables a better estimation of the fig- ure of merit for the quality of superconducting nanowires: the constriction factor $C$ [7], ratio of the switching and depairing currents ($C = I_{\text{sw}}/I_{\text{dep}}$), since reaching higher fractions of the depairing current gives rise to higher in- ternal detection efficiency and lower intrinsic jitter [6].
The kinetic inductance dependence on bias current is determined by measuring the self-resonance of a su- perconducting nanowire in a coplanar waveguide (CPW) structure, using a vector network analyzer (VNA). The resonances were measured in both transmission and re- flection modes by analyzing the complex spectral re- sponse. Measurement of the self-resonance has been demonstrated for meandered nanowires [24], however, the change of the kinetic inductance at the highest achievable bias current relative to the zero bias current case was less than 10%, making it difficult to distinguish whether the experiment falls within the fast or slow relaxation cat- egory and giving rise to significantly different depairing current predictions [23]. Here we demonstrate a kinetic- inductance change as high as 31% for tungsten silicide (WSi) and 28% for niobium nitride (NbN) nanowires, which allows us to conclude that the experiment falls into the fast relaxation regime by comparing the qual- ity of the fit of the two models. The improvement could be attributed to several factors such as optimized mate- rial [25], lower base temperature and use of a cryogenic
| |
190408596/3
|
## EXPERIMENTAL SETUP
The experimental setup is illustrated in Fig. 2. We measured the resonant frequency of the SNSPD-like res- onators (Fig. 1) both in transmission mode [24] and in reflection mode (Fig. 2). The devices were cooled to a base temperature of 1.05 K with a cryocooler composed of a pulse tube followed by a Helium-4 sorption cooler.
The device resonance was measured with a 300 kHz- 6 GHz VNA. The output signal from the VNA was atten- uated by 20 dB at both the 40 K and 4 K stages, before entering the input port of a 20 dB directional coupler. The transmission port of the coupler was 50 $\Omega$ termi- nated, while the coupling port, was connected to an RF switch on the 1 K stage. One port of the switch was connected to a through device, which consisted of a su- perconducting CPW used for calibration purposes. The RF switch was used to achieve the same electrical envi- ronment between the calibration device and the device under test.
While measuring in the transmission mode, the out- put port of the resonator CPW, was connected to a sec- ond switch followed by a cryogenic bias tee. The DC port was used to current bias the nanowire, while the RF port was fed to the input of a SiGe cryogenic amplifier (Cosmic Microwave, CITLF1 [30]). The bias tee and am- plifier were mounted and thermalized to the 4 K stage. The amplified RF signal was fed to the input port of the VNA. Finally, the isolated port of the directional coupler was 50 $\Omega$ terminated to guarantee current flow through the nanowire. In reflection mode, the nanowire was con- nected to the coupler on one side and grounded on the other. In this configuration, the isolation port of the di- rectional coupler connected the nanowire to the bias tee and amplifier. For both scenarios, the power output of the VNA was adjusted such that the RMS current flowing through the resonator was of the order of 100 nA, which is small to prevent a shift in the resonant frequency.
## MODELS
The measured resonance peaks were fitted using a RLC resonator model with a purely reactive bypass chan- nel. For the reflection mode measurement, the resonance was fitted using a double notch filter at the resonant fre- quency. The magnitude and phase functions of $S_{11}(\omega)$ are written as
\begin{subequations}
$$
\begin{aligned}
\|S_{11}(\omega)\| &= - I\frac{\big(\frac{1}{2}\Gamma\big)^2}{(\omega - \omega_r)^2 + \big(\frac{1}{2}\Gamma \big) ^2} \text{,} \\
\text{arg} \big\{S_{11}(\omega) \big\} &= -180 + 2\times \tan^{-1}\big[2Q (1 - \frac{\omega}{\omega_r}) \big] \text{,}
\end{aligned}
$$
\end{subequations}
where $\Gamma$ is the full-width at half-maximum of the Lorentzian function, $I$ is the peak height, $\omega_r$ is the reso- nant frequency and $Q$ is the quality factor.
For the transmission mode measurement, the reso- nance is still been modeled as a Lorentzian function, but accounts for the effect of a bypass channel, modeled as a pure capacitance, in a correction factor. We define $S_{21}(\omega)$ according to
$$
\|S_{21}(\omega)\| = I\frac{\big(\frac{1}{2}\Gamma\big)^2}{(\omega - \omega_r)^2 + \big(\frac{1}{2}\Gamma \big) ^2} \Big| 1 - \xi (\omega - \omega_r) \Big| ^2 \text{,}
$$
where the correction factor to the Lorentzian function in (2) is valid for purely reactive bypass channels and $\xi$ is a constant representing the coupling between the res- onator and the reactive channel. For further information regarding the physical meaning of $\xi$, we direct readers to the supplementary information of Weinstein and Schwab [31].
Once the resonant frequency of the nanowire was evaluated, we could estimate the change in kinetic in- ductance with increasing bias current according to $\omega_r \propto 1/\sqrt{LC}$ for an RLC resonator. We then fitted the kinetic inductance ratios as obtained using the two relaxation models from Clem and Kogan [23]
\begin{subequations}
$$
\begin{aligned}
y_{\text{fr}}(x) &= (1 - x^n)^{1/n} \text{,}
\\
y_{\text{sr}}(x) &= y_0 - (y_0 - 1)(1 - x^n)^{1/n} \text{,}
\end{aligned}
$$
\end{subequations}
where $y = \mathcal{L}_{\text{k}}(q,t)/\mathcal{L}_{\text{k,0}}(t)$ is the ratio between the ki- netic inductance of the biased superconducting nanowire and the kinetic inductance at zero bias current, $y_0$ and $n$ are fixed parameters defined by Clem and Kogan [23] for specific temperature ratios $t = T/T_c$, $x = |j_{\text{s}}| /j_{\text{d}}(t)$ is the ratio between the bias current density and the depairing current density and the subscripts “ fr” and “ sr” stand for “ fast relaxation” and “ slow relaxation” respectively. The difference between the two models is related to the char- acteristic timescale of variation of $j_{\text{s}}$, the current-biased experiment characteristic time $\tau_{\text{exp}}$, with respect to the relaxation time of the superconductor ($\tau_{\text{s}}$). We refer to fast relaxation if the experimental time constant is much larger than the characteristic superconductor relaxation time, while for slow relaxation, the experimental time constant is much smaller. The accuracy of the fitting functions (3a) and (3b) compared to the full numerical solution presented in [23] is 1% for the fast relaxation model and 0.5% for the slow relaxation model. As a com- parison, we also calculate the depairing current using a fit to the full numerical results of the fast relaxation model using the approach of [23] and keeping 15000 modes in the numerical calculations. The numerical results pro- vide a better match to the experimental results than the approximate equation with only a small change in the extracted depairing current when compared to the ap- proximate fit of (3a). Within both models, the depairing current density $j_{\text{d}}(T)$ is the only fitting parameter.
| |
231002763/17
|
## SUMMARY AND CONCLUSIONS
random location on the line of sight subsequently miss the Earth. Consider however neutrinos which are not emitted into the solid angle of the Earth which subsequently decay. It may happen that one of the secondaries arising at some angle w.r.t. its direction of flight, enters the solid angle of the detector on Earth. Such neutrinos can still have very high energies but travel a longer distance (time) and can drop off from the coincidence with the electromagnetic component. Assessing whether this is a significant effect requires however a dedicated study of differential distributions $d \Gamma /d \Phi$ (2).
## Summary and conclusions
We have analysed the decays of high and ultra-high energy neutrinos under an assumption that these particles are spacelike – a hypothesis already standing for some time, inspired by theoretical and experimental results. It has been shown that not observing ultra-high energy neutrinos on Earth can be explained as due to decays, $\nu_{\alpha}\rightarrow \nu_{\alpha} \; \nu_{\beta} \bar{\nu}_{\beta}$, if neutrinos were spacelike fermions. We have calculated the Standard Model width for such decays, with the tachyonic neutrino mass being the only quantity not precisely known. Subsequently we derived an exact expression for survival probability on the way to Earth, parametrised by the emission energy, $E_e$, or energy measured on Earth, $E_d$, and the redshift of the source, $z_e$, in the environment of expanding Universe.
We exploited the fact of observing five high neutrino events ($i$) – ($v$) to discuss and demonstrate a possibility to set approximate upper limits on the tachyonic neutrino mass. We have shown that recording the highest energy 13 PeV event allows to put an upper mass limit of about $0.6\mbox{div} 0.8$ eV, subject of reasonable assumptions. These approximate values fall very near to the range allowed by the latest measurement of KATRIN, $\kappa <$ 0.55 eV at 90% c.l. The consistency of our cosmology-based estimations with the results from a terrestrial $\beta$ -decay experiment is remarkable. We also show that a neutrino mass in the approximate range indicated in Fig. 6 can explain the non-observation of ultra- high energy tachyonic neutrinos as due to their decays, provided the neutrino mass is not smaller than about 0.15 eV. If the fact of ”non-observation” was confirmed experimentally on a sufficient statistics, this could lead to establishing a lower limit on the neutrino mass – first result of this kind as regards the mass of the neutrino. Lastly, the smallness of the neutrino masses seems to provide some rationale to the hypothesis of tachyonic neutrinos. If the neutrino mass amounted to only a few eV, neutrino fluxes from nearby galaxies would be depleted at already TeV energies.
This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.
The authors declare that they have no known competing financial interests or personal relation- ships that could have appeared to influence the work reported in this paper.
No data was used for the research described in the article.
| |
231002763/5
|
## DECAYS OF SPACELIKE NEUTRINOS
Figure1: Feynmann diagram of the tachyonic neutrino/antineutrino three-body decay. The neutrino momenta and tachyonic masses that appear in Eq. 3 are indicated in brackets.
\begin{multline} |M|^2 = \frac{64}{\sqrt{(uk)^2+\kappa^2}\sqrt{(up)^2+\kappa^2} \sqrt{(ul)^2+\mu^2} \sqrt{(ur)^2+\mu^2}}\\
\Bigg (\kappa^4\mu^4 + \kappa^4\Big[\mu^2 \Big((ul)^2+(ur)^2\Big) +(ul)^2(ur)^2\Big] + \mu^4\Big[\kappa^2 \Big((up)^2+(uk)^2\Big) +(up)^2(uk)^2\Big] \\
+ \kappa^2\mu^2\Big[(kr) \Big((uk)(ur)-(up)(ul)\Big) +(up)^2(ur)^2 + (uk)^2(ul)^2\Big]\\
+ (kr) \Big((uk)(ul)-(up)(ur)\Big) \Big(\kappa^2 (ul)(ur)- \mu^2(uk)(up)\Big) - (uk) (up) (ur) (ul) (kr)^2 \Bigg), \end{multline}
where $u$, as above, denotes the four-velocity of the preferred frame (CNB) and the scalar products are of the form $uk=u^{\lambda}k_{\lambda}$ etc. Since below we calculate the width for the decay in the preferred frame (CNB), we put $u=(1,0,0,0)$, for which $uk=k^0=E$. In view of the complexity of these calculations, the width $\Gamma$ cannot be easily derived as a closed form expression. Instead, we have developed a simple effective formula on the basis of dimensional and numerical analysis, accurate to the 8-th decimal in the energy range under study (TeV-PeV)
$$
\Gamma (E,\kappa,\mu) = f\; n_f\, G_F^2 (\kappa^4 + 4 \mu^4)E,
$$
where $f=\frac{5}{18 (2\pi)^3}$ and $n_f=3$ accounts for the three neutrino flavours in the final state of the $\nu\bar{\nu}$ pair (cf. Fig. 1). Understandably, the powers of the masses and energy, integers to a very high precision, add up to five. We note the strong dependence of the width on the neutrino masses as well as the dominating weight of the mass $\mu$. The numerical calculations require a very high working precision to obtain a numerically stable result.
Given the mass differences squared measured in oscillation experiments, one can derive the masses of the heavier mass states as a function of the mass of the lowest mass state, $m_1$: $m_2(m_1)=\sqrt{\Delta m_{21}^2 + m_1^2}$ and $m_3(m_1)=\sqrt{\Delta m_{21}^2 + \Delta m_{32}^2 + m_1^2}$, where $\Delta m_{21}^2 = 7.4\times 10^{-5}$ eV $^2$ and $\Delta m_{32}^2 = 2.5\times 10^{-3}$
| |
231002763/7
|
## DECAYS OF SPACELIKE NEUTRINOS
Defining the four-momentum, $k^{\mu}$, in the standard way as
$$
k^{\mu}=\kappa \,dx^{\mu}/d\lambda,
$$
where $\kappa$ is the particle mass, we obtain the dispersion relation in the form
$$
g_{\mu\nu} k^{\mu}k^{\nu} = -\kappa^2.
$$
In the case of a flat FLRW one has
$$
(k^0)^2 - a(t)^2 \Big ( (k^r)^2 + r^2 \left((k^\vartheta)^2 + (k^\varphi)^2 \sin^2\vartheta \right) \Big) = -\kappa^2
$$
so taking into account that
$$
g_{ij} k^i k^j= a(t)^2 \Big ( (k^r)^2 + r^2 \left((k^\vartheta)^2 + (k^\varphi)^2 \sin^2\vartheta \right) \Big) \equiv \boldsymbol{m}{k}^2
$$
is the momentum squared, we can rewrite (9) in the standard form
$$
(k^0)^2 - \boldsymbol{m}{k}^2 = -\kappa^2.
$$
Now, we can simplify the considerations by taking into account the fact that the direction of the spacelike neutrino is unchanged throughout its motion. This implies $d\vartheta d\varphi=0$ and the particle line element (6) reduces to the form
$$
-d\lambda^2= dt^2-a(t)^2 dr^2.
$$
In such a case $k^{\vartheta}=k^{\phi}=0$ and $\boldsymbol{m}{k}^2=a(t)^2 (k^r)^2$. Substituting () into the geodesic equations
$$
\frac{d^2 x^{\mu}} {d\lambda^2} + \Gamma^{\mu}_{\alpha\beta} \frac{d x^{\alpha}}{d\lambda} \frac{d x^{\beta}}{d\lambda}=0
$$
yields straightforwardly
$$
\kappa \frac{d k^{\mu}} {d\lambda} + \Gamma^{\mu}_{\alpha\beta} k^{\alpha}k^{\beta}=0
$$
and by calculating the connection coefficients for this case we can reduce (14) to only one independent equation of the form
$$
k^0 dk^0 + |\boldsymbol{m}{k}| \frac{da}{a}=0.
$$
Therefore the geodesic motion of the spacelike neutrino in a flat FLRW spacetime is determined by the dispersion relation (11) and the geodesic condition (15). Solving the system of these two equations one obtains
$$
\begin{aligned}
(k^0)^2 &=& C^2 a^{-2} -\kappa^2\\
|\boldsymbol{m}{k}| &=& Ca^{-1}.
\end{aligned}
$$
$$
\begin{aligned}
(k^0)^2 &=& C^2 a^{-2} -\kappa^2\\
|\boldsymbol{m}{k}| &=& Ca^{-1}.
\end{aligned}
$$
| |
231002763/12
|
## RESULTS AND DISCUSSION
Contours of fixed neutrino survival probability $p_{\rm c}=0.1$, for $E_d=6.3$ PeV (Glashow resonance), 13 PeV (the highest energy event of IceCube) and 30 PeV for comparison.
| |
231002763/14
|
## RESULTS AND DISCUSSION
Survival probability as a function of tachyonic neutrino energy for $z_e=1.839$ (PKS 1502+106). Black dashed curve shows the prediction for the KATRIN limit of $\kappa=0.55$, solid curves are drawn for $\kappa$ =0.35, 0.15 and 0.05 eV. Red vertical dashed lines mark energies of 300 TeV, 6.3 PeV and 13 PeV. The shaded area marks the allowed region for the tachyonic neutrino mass, between the KATRIN limit and the minimal value.
| |
231002763/2
|
## WHY SPACELIKE NEUTRINOS
interacting with photons of the cosmic microwave background radiation [0, 0] but can also be ex- plained as a consequence of a maximal energy reachable in cosmic sources. Since UHE neutrinos are expected to carry $3 \mbox{div} 5$% of the primary hadron energy, they should be observed on Earth with energies even up to a few EeV [0,0,0,0]. However this is not the case so the question of not observing UHE neutrinos is still open. Lorentz symmetry violation has been considered in this context [0, 0, 0], involving a kinematical high-energy cutoff, possibly leading to non-observation of neutrinos beyond a certain energy. According to another concept, UHE primary and secondary charged particles, spiraling in a magnetic field in the source, are subject to a significant radiative energy loss prior to eventually decaying into neutrinos. Also sensitivity of the present detectors to neutrinos in the UHE range can be insufficient.
In the present paper we adopt a hypothesis that neutrinos are spacelike (thus superluminal) fermions and consider kinematically allowed neutrino decays to explain the reduction of the neutrino flux in the high energy end of the spectrum. A tachyonic particle is characterised by a spacelike energy-momentum dispersion relation, $E^2 - \boldsymbol{m}{k}^2 = -\kappa ^2$, where $\kappa$ denotes the tachyonic mass, in contrast to the inertial mass, $m$, satisfying the relation $E^2 - \boldsymbol{m}{p}^2 = m ^2$ for massive particles. We derive the neutrino decay width within the framework of a slightly modified Standard Model in the neutrino sector and translate it into the decay probability in the expanding Universe. We determine upper limits on the tachyonic neutrino mass, $\kappa$, from the observation on Earth of the highest energy cosmological neutrino events and find these results very close to the independent measurement of $m_{\nu}^2$ in the tritium decay experiment KATRIN. In the following the term ”neutrino” implicitly refers to the ”spacelike neutrino” unless stated otherwise.
## Why spacelike neutrinos
Neutrinos have been associated with superluminality since decades. The initial underlying justifi- cation consisted in repeated occurrences of negative or consistent with negative central values for the electron antineutrino mass squared, observed in numerous tritium decay experiments, to quote only the most recent: $m^2= -0.6 \pm 2.2{(\rm stat.)} \pm 2.1{\rm (syst.)}$ [0] (Mainz Collaboration, 2005) and $m^2= -0.67 \pm 2.53{(\rm tot.)}$ [0] (Troitsk Collaboration, 2011). These results were superseded by those obtained in the presently running experiment KATRIN which achieved an unprecedented accuracy. Their first measurement again yielded a negative central value of the mass squared, $m_{\nu}^2=-1.0^{+0.9}_{-1.1}$ eV $^2$ [0] (2019), while the following measurement period ended with a positive central value, $m_{\nu}^2=0.26\pm 0.34$ eV $^2$ (2022), however consistent with being negative within even less than $1\sigma$. In contrast to any other elementary object, letting aside the common prejudice, one must accept that according to the present evidence, the four-momentum of the neutrino may as well be spacelike. Admitting the above requires providing an adequate theoretical description of spacelike neutrinos at the quantum field theory level.
Early attempts to describe tachyonic neutrinos within the standard (Einsteinian) relativity were unsuccessful at any level and did not lead to a solution of the essential problems within this frame-
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.