Scoto-seesaw-DM-pheno

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

The simplest scoto-seesaw model:

WIMP dark matter phenomenology and Higgs vacuum stability

Sanjoy Mandal,1, ∗ Rahul Srivastava,2, † and José W. F. Valle1, ‡


1 AHEP Group, Institut de Fı́sica Corpuscular –
CSIC/Universitat de València, Parc Cientı́fic de Paterna.
C/ Catedrático José Beltrán, 2 E-46980 Paterna (Valencia) - SPAIN
2 Department of Physics, Indian Institute of Science Education and Research - Bhopal,
Bhopal Bypass Road, Bhauri, Bhopal 462066, India
arXiv:2104.13401v2 [hep-ph] 22 Jun 2021

We analyze the consistency of electroweak breaking, neutrino and dark matter


phenomenology within the simplest scoto-seesaw model. By adding the minimal dark
sector to the simplest “missing partner” type-I seesaw one has a physical picture for
the neutrino oscillation lengths: the “atmospheric” mass scale arises from the tree-
level seesaw, while the “solar” scale is induced radiatively, mediated by the dark
sector. We identify parameter regions consistent with theoretical constraints, as
well as dark matter relic abundance and direct detection searches. Using two-loop
renormalization group equations we explore the stability of the vacuum and the
consistency of the underlying dark parity symmetry. One also has a lower bound for
the neutrinoless double beta decay amplitude.

1. INTRODUCTION

The discovery of neutrino oscillations [1, 2] implies that at least two neutrinos are massive.
There has by now been strong evidence, at different scales, for the existence of cosmological
dark matter, the basic understanding and interpretation of which we also lack [3]. The main
current neutrino mass generation paradigms are the seesaw and the scotogenic mechanism,
which also accounts for dark matter as the mediator of neutrino mass, as a result of an
assumed Z2 symmetry. Both mechanisms give mass “democratically” to all neutrino states,
according to the structure of the relevant Yukawa couplings.
The simplest “scoto-seesaw” extension of the Standard Model [4] combines these two
main paradigms within its minimal SU(3)c ⊗ SU(2)L ⊗ U(1)Y framework. In such hybrid

[email protected]

[email protected]

[email protected]
2

scenario the atmospheric scale comes from the tree level seesaw, while the solar scale is
mediated by the radiative exchange of dark states, i.e.
 2 2 2 !2
2

v 1 λ 5 v
∆m2ATM = Y2 , ∆m2SOL ≈ Mf Y2f , (1)
2MN N 32π 2 Mf2 − m2η
where MN is the “right-handed” neutrino mass, Mf and mη are “dark-sector” masses and
Y2N , Y2f are corresponding Yukawa coupling strengths.
One sees that the solar splitting will be non-zero as long as λ5 6= 0. Moreover, one
accounts naturally for the hierarchy between the solar and atmospheric scales observed in
the experimental data [5]. The corresponding ratio of squared solar-to-atmospheric mass
splittings for normal and inverted mass hierarchy are found to be [5, 6]:
∆m2SOL +0.0027 ∆m2SOL
NO: = 0.0294 −0.0023 , IO: = 0.0306+0.0028
−0.0025 . (2)
∆m2ATM ∆m2ATM
Altogether, the interplay of the “seesaw” and “dark-sectors” provide an interesting way to
describe lepton number violation and neutrino mass generation. Indeed, the scoto-seesaw
model has a viable weakly interacting massive particle (WIMP) dark matter candidate and
accounts for the observed neutrino masses, including the solar-to-atmospheric hierarchy.
The aim of this work is to explore the scoto-seesaw model in more detail.
The paper is organized as follows. In section 2 we briefly describe the model, giving the
details of the new fields and their interactions. In section 3, we describe the tree level and
radiative neutrino mass generation. In section 4 we study the parameter space for the case of
a scalar dark matter candidate. In section 6 we look at vacuum stability in the scoto-seesaw
model and show that, over large parameter regions, the vacuum is stable all the way up to
the Planck scale. In section 7 we examine the robustness of the dark parity symmetry under
renormalization group (RG) evolution of the parameters. We finally conclude in section 8.

2. MINIMAL SCOTO-SEESAW MODEL

The minimal combination of the seesaw mechanism and the scotogenic model was pro-
posed in Ref. [4] 1 . It clones the simplest “missing partner” (3,1) version of the Standard
Model seesaw mechanism suggested in [8, 9] with the minimal scotogenic model proposed
in [10]. We now describe in detail both the fermionic and the scalar sectors of the model.

2.1. The Yukawa Sector

We now briefly recall the basic features of the minimal scoto-seesaw model [4]. The new
particles and their charges are given in Table I, where the family index a runs from 1 to 3.
1
One can have scoto-seesaw realizations based on (3,2) seesaw extensions [7]. While they have new inter-
esting features, one looses the interesting prediction in Eq. (1).
3

Standard Model New Fermions New Scalar


La ea H N f η
SU (2)L 2 1 2 1 1 2
U (1)Y -1/2 -1 1/2 0 0 1/2
Z2 + + + + − −

TABLE I: Matter content and charge assignment of the minimal scoto-seesaw model.

In Table. I the additional Z2 symmetry is the “dark parity” responsible for the stablity
of the dark matter candidate. All the Standard Model particles and N are even under this
dark Z2 parity, while the dark sector, consisting of one fermion f and one scalar η, is odd
under Z2 .
The full Yukawa sector can be split as

L = LSM + LATM + LDM,SOL (3)

where LSM is the Standard Model Lagrangian, while


1
LATM = −YNa L̄a H̃N + MN N c N + h.c, (4)
2
induces the type-I seesaw neutrino mass (atmospheric neutrino mass scale) after the elec-
troweak symmetry breaking. Note also that throughtout this work repeated indices imply
summation, with H̃ = iσ2 H ∗ , σ2 being the second Pauli matrix.
The Lagrangian responsible for the solar and dark sector is given by
1
LDM,SOL = Yfa L̄a η̃ f + Mf f c f + h.c. (5)
2
It induces the solar neutrino mass scale as discussed in Sec. 3.

2.2. The Scalar Sector

Apart from the Standard Model (SM) Higgs doublet H we have a scalar doublet η carrying
the same quantum numbers, but with Z2 -odd parity. The SU(3)c ⊗ SU(2)L ⊗ U(1)Y gauge
invariant scalar potential is given by

V = −µ2H H † H + m2η η † η + λ(H † H)2 + λη (η † η)2 + λ3 (H † H)(η † η) + λ4 (H † η)(η † H)


λ5
(H † η)2 + h.c.

+ (6)
2
We now turn to the consistency conditions of the potential. The following restrictions
must hold so as to ensure that the scalar potential is bounded from below and has a stable
4

vacuum at any given energy scale µ:

λ(µ) > 0, λη (µ) > 0, (7)

q
λA ≡ λ3 (µ) + 4λ(µ)λη (µ) > 0, (8)

q
λB ≡ λ3 (µ) + λ4 (µ) + 4λ(µ)λη (µ) − |λ5 (µ)| > 0. (9)

where λi (µ) are the values of the quartic couplings at the running scale µ. In order to have
an absolutely stable vacuum, one must satisfy the conditions given in Eqs. (7), (8) and (9)
at each and every energy scale. To ensure perturbativity, we take a conservative approach
of simply requiring that the scalar quartic couplings in Eq. (6) obey ≤ 4π.

2.3. Mass Spectrum

In order to ensure dark matter stability the Z2 symmetry should remain unbroken. This
means that the Z2 odd scalar η should not acquire a nonzero vacuum expectation value
(VEV). As a result, electroweak symmetry breaking is driven simply by the VEV of H. The
fields η and H can be expanded as follows
! !
H+ η+
H= √ , η= √ (10)
(v + h + iφ0 )/ 2 (η R + iη I )/ 2

Exact conservation of the Z2 symmetry forbids the mixing between the Higgs and the dark
doublet η. The components of η have the following masses

1
m2ηR = m2η + (λ3 + λ4 + λ5 ) v 2 (11)
2
1
m2ηI = mη + (λ3 + λ4 − λ5 ) v 2
2
(12)
2
1
m2η+ 2
= mη + λ3 v 2 . (13)
2

The difference m2ηR − m2ηI depends only on the parameter λ5 which, we will show later,
is also responsible for smallness of solar neutrino mass scale. The conservation of the Z2
symmetry also makes the lightest of the two eigenstates η R and η I a viable scalar dark
matter candidate2 , as we explore in what follows.

2
Throughout this work we assume that the dark fermion f is heavier than the dark scalars η R and η I .
5

3. NEUTRINO MASSES

At tree-level this model gives rise to the following neutrino mass matrix Mab
ν ,
 Y 1v 
0 0 0 √N2
 Y 2v 
ab
 0 0 0 √N2 
Mν =   0

YN3 v  (14)
 0 0 √
2 
YN1 v YN2 v YN3 v

2

2

2
MN

in the basis (La , N )T . Notice the (3,1) structure of the seesaw [8, 9], as a result of which one
sees that the N pairs off with one combination of the doublets in La through their Dirac-like
couplings.
This clearly leads to a projective structure for the tree-level neutrino mass matrix
v2
MνTREE = − Y aY b , (15)
2MN N N
where a, b = 1, 2, 3 are family indices of the lepton doublets. One sees from Eq. (15) that
for “sizeable” Yukawa couplings, YN ∼ O(1), in order to reproduce the required value of
the atmospheric scale, heavy neutrinos must lie at mass scale MN ∼ O(1014 GeV). Smaller
values of the Yukawa coupling YN would require correspondingly lower seesaw scale MN .
The solar mass scale arises from Fig. 1 involving the exchange of the scalar and fermionic
dark mediators η and f . Although these corrections also have a projective structure, they

hHi hHi

η η

L L
f

FIG. 1: Solar neutrino mass scale induced from radiative dark-sector exchange.

break the “missing-partner” nature of the (3,1) type-I seesaw mechanism3 . The total neu-
trino mass has the form
v2
Mab
νTOT = − Y a Y b + F(mηR , mηI , Mf )Mf Yfa Yfb , (16)
2MN N N
3
The situation is analogous to neutrino mass generation in bilinear broken R-parity supersymmetry [11–13].
6

where the first term is the tree-level seesaw part, and the loop function F characterizes the
quantum correction arising from Fig. 1. This is responsible for inducing the solar mass scale.
The loop function F is expressed as the difference of two B0 -Veltman functions, namely,
    
1  m2ηR log Mf2 /m2ηR m2ηI log Mf2 /m2ηI
F(mηR , mηI , Mf ) = −  (17)
32π 2 Mf2 − m2ηR Mf2 − m2ηI

Since both terms in Eq. (16) have a projective nature, one out of the three neutrinos remains
massless.
From the eigenvalues of the neutrino mass matrix Mab
ν one can estimate the atmospheric
and solar square mass differences as,

2 2 !2
v2 2 λ5 v 2
 
1
∆m2ATM = Y , ∆m2SOL ≈ Mf Y2f . (18)
2MN N 32π 2 Mf2 − m2ηR

where we take Mf2 , m2ηR , Mf2 − m2ηR  λ5 v 2 and Y2` = (Y`e )2 + (Y`µ )2 + (Y`τ )2 for ` = N, f .
It follows that the ratio between the solar and atmospheric square mass differences can be
written as:
2 !2  2
∆m2SOL Y2f

1 MN Mf
≈ λ5 2 (19)
∆m2ATM 16π 2 Mf − m2ηR Y2N

From Eq. (18) it is clear that one can fit the observed atmospheric and solar mass square
differences in many ways as long as one takes an adequately small value for λ5 . Moreover,
Eq. (19) nicely reproduces Eq. (2). In the following we list some choices which can satisfy
both the solar and atmospheric scales, as well as have η R as the scalar WIMP dark matter:

• MN ∼ 1014 GeV, Mf ∼ 1012 GeV, mηR ∼ 103 GeV, YN ∼ 0.4, Yf ∼ 0.4,

• MN ∼ 1012 GeV, Mf ∼ 104 GeV, mηR ∼ 103 GeV, YN ∼ 0.1, Yf ∼ 10−4 ,

• MN ∼ 1014 GeV, Mf ∼ 105 GeV, mηR ∼ 103 GeV, YN ∼ 0.4, Yf ∼ 10−4 ,

• MN ∼ 106 GeV, Mf ∼ 106 GeV, mηR ∼ 103 GeV, YN ∼ 10−5 , Yf ∼ 10−4 .

The upshot of this discussion is that one can easily fit the solar and atmospheric scales for
reasonable parameter choices. For example, for sufficiently small λ5 values, one can choose
I
a reasonable Yukawa coupling Yf and Mf > ∼ O(TeV). In section 4 we show that either η or
η R can, indeed, be taken as a consistent WIMP dark matter candidate.
7

4. PHENOMENOLOGY OF SCALAR WIMP DARK MATTER

In this section we collect the results of our analysis of dark matter phenomenology. In
addition to ensuring radiative generation of neutrino masses, the Z2 symmetry in the dark
sector ensures the stability of “lightest dark particle” (LDP). Such LDP is in principle a
viable dark matter candidate. There are three LDP options. The first is the dark fermion
f . The others are the real and imaginary parts of the η scalar, η R and η I . In our analysis,
we assume scalar dark matter, with the condition λ5 < 0 on the quartic coupling λ5 . As a
result η R will be our dark matter candidate (the opposite scenario with λ5 > 0 would have
η I as the dark matter particle).
In order to calculate all the vertices, mass matrices, tadpole equations etc the model is
implemented in the SARAH package [14]. On the other hand, the thermal component of
the dark matter relic abundance, as well as the dark matter-nucleon scattering cross section,
are determined using micrOMEGAS-5.0.8 [15].

4.1. Relic density

As shown in Fig. 9 (Appendix A), there are several dark matter annihilation and coan-
nihilation diagrams present in the scoto-seesaw model. They involve annilation to quarks
and leptons, SM gauge bosons and the Higgs boson. Altogether, they determine the relic
abundance of our assumed LDP, η R . Our numerical scan is performed varying the input
parameters as given in Table II.

Parameters Range
m2η [1002 , 50002 ] (GeV2 )
λ3 [10−5 , 1]
λ4 [10−5 , 1]
|λ5 | [10−5 , 10−3 ]

TABLE II: Ranges of variation of the input parameters used in our numerical scan.

In Fig. 2 we show the relic density as a function of the mass of the scalar dark matter
candidate η R . The narrow horizontal band is the 3σ range for cold dark matter derived from
the Planck satellite data [16]:

0.1126 ≤ ΩηR h2 ≤ 0.1246. (20)

Only for solutions falling exactly within this band the totality of the dark matter can be
explained by η R . The relic density for the cyan points in Fig. 2 lies within the above 3σ
range, whereas the relic density for blue and gray points is above and below the 3σ range.
8

FIG. 2: Relic abundance as a function of the dark matter mass mηR . Cyan points inside the black lines
fall within the measured 3σ cold dark matter relic density range given by Planck satellite data, Eq. (20).
Gray (blue) points outside the narrow band give under (over) abundance of dark matter, respectively.

One sees from Fig. 2 that the correct relic density can be obtained in three mass ranges:
mηR < 50 GeV, 70 GeV < mηR < 100 GeV and mηR > 550 GeV. The reasons for these
mass gaps can be understood by looking in detail into the η R annihilation channels (see
Appendix A). The first dip occurs at mηR ∼ MZ /2 and corresponds to annihilation via s-
channel Z exchange. The second depletion of the relic density happens around mηR ∼ mh /2
and corresponds to annihilations via s-channel Higgs boson exchange. This becomes very
efficient when the SM-like Higgs h is on-shell, precluding us from obtaining a relic density
matching Planck observations. Notice that the second dip is more efficient than the first one,
as the Z-mediated dip is momentum suppressed. For heavier η R masses, quartic interactions
R + −
with gauge bosons become effective. For mηR > ∼ 80 GeV, annihilations of η into W W
and ZZ via quartic couplings are particularly important, thus explaining the third drop
in the relic abundance. In the mass range mηR ≥ 120 GeV, η R can annihilate also into
two Higgs bosons, hh. When mηR ≥ mt , a new channel η R η R → tt̄ opens up. All these
annihilation channels make dark matter annihilation very efficient, and it is difficult to
obtain the correct relic density. For very heavy mηR the relic density increases due to
the suppressed annihilation cross section, which drops as ∼ m12 . Notice also that the
ηR
9

coannihilation channels with η I and η ± may occur in all regions of the parameter space,
with the effect of lowering the relic dark matter density.

4.2. Direct detection

Let us now study the direct detection prospects of our dark matter η R . In our model,
the tree-level spin-independent η R -nucleon cross section is mediated by the Higgs and the Z
portals, see Fig. 3. Notice that, as the η doublet has non-zero hypercharge, the η R -nucleon

ηR ηR ηR ηI

h Z

q q q q

FIG. 3: Higgs and Z-mediated tree-level Feynman diagrams contributing to the elastic
scattering of η R off nuclei.

spin-independent (SI) cross-section is mediated by the Z-boson. Generally this exceeds the
current limit from direct detection experiments like XENON1T [17]. However this can be
easily avoided by taking non-zero λ5 . In this case there is a small mass splitting between
η I and η R , so that the interaction through the Z-boson is kinematically forbiden or leads
to inelastic scattering. As a result, for nonzero λ5 , the η R -nucleon interaction via the Higgs
will be the dominant one. The coupling between η R and the Higgs boson depends on
λ345 = λ3 + λ4 + λ5 and the η R -nucleon cross section is given by

λ2345 m4N fN2


σ SI = (21)
4πm4h (mηR + mN )2

where mh is the mass of SM Higgs boson and mN is the nucleon mass, i.e. the average of the
proton and neutron masses. Here fN is the form factor, which depends on hadronic matrix
elements. In Fig. 4 we show the spin-independent η R -nucleon cross section as a function
of the η R mass, for the range of parameters covered by our scan given in Table II. The
color code in Fig. 4 is the same as in Fig. 2. The red line denotes the latest upper bound
from the XENON1T collaboration. There are constraints from other experiments as well,
such as LUX [25] and PandaX-II [26], but weaker when compared to the XENON1T limit.
We also show the projected sensitivities for the PandaX-4t [18], LUX-ZEPLIN(LZ) [19],
10

FIG. 4: Spin-independent WIMP-nucleon elastic scattering cross section versus the dark matter mass
mηR , with the same colour code as in Fig. 2. The solid red line denotes the recent upper bound from the
XENON1T experiment [17]. A broad region of experimentally viable scoto-seesaw model points in cyan
fall within future projected sensitivities for the PandaX-4t [18] (green), LZ [19] (dashed purple),
XENONnT with 20 ton-yr exposure [20] (blue), DarkSide-20k [21] (magenta), DARWIN [22] (black) and
ARGO [23] (brown) proposals. The dot-dashed orange line corresponds to the “neutrino floor” coming
from coherent elastic neutrino scattering [24]. The upper triangular white region violates perturbativity.

XENONnT [20], DarkSide-20k [21], DARWIN [22] and ARGO [23] experiments. The lower
limit corresponding to the “neutrino floor” from coherent elastic neutrino scattering is also
indicated. We see from Fig. 4 that there are low-mass solutions with the correct dark matter
relic density. However, most of these are ruled out by the XENON1T direct detection cross
section upper limits. Moreover, there are also tight constraints on low mass dark matter
from collider searches, as we will discuss in the next section.
11

5. COLLIDER CONSTRAINTS

In this section we confront our scalar dark matter candidate η R with the latest data from
particle colliders, in particular the LHC. First we note that, if η R /η I are light enough, there
are two additional decay channels for the SM-like Higgs boson,
s
R R
2 2
v λ345 4m2ηR
Γ(h → η η ) = 1− (22)
32πmh m2h
s
2 2 λ345 2 2
(mηI − mηR + 2 v ) 4m2ηI
Γ(h → η I η I ) = 1 − (23)
8πv 2 mh m2h
Note that due to LEP limit mη± > mW [27], there is no phase space for the two body decay
h → η ± η ± . The decay mode Γ(h → η R η R ) contributes to the invisible Higgs decay width,
constrained by the LHC experiments, e.g. the CMS experiment [28],

BR(h → Inv) ≤ 0.19.

The SM-like Higgs boson h also couples to the charged Higgs η ± , contributing to the diphoton
decay channel h → γγ 4 . To quantify the deviation from the Standard Model prediction, we
define the following parameter
BR(h → γγ)
Rγγ = . (24)
BR(h → γγ)SM
The value we use for the Standard Model is BR(h → γγ)SM ≈ 2.27 × 10−3 .

FIG. 5: Invisible Higgs branching ratio (left panel) and Rγγ (right panel) as a function of
the dark matter mass mηR . The color code is same as in Fig. 2. The shaded red region in
the left panel is excluded from the LHC constraint on the invisible Higgs decay [28], while
the shaded green band in the right is allowed by ATLAS measurements of Rγγ [33].

4
Note that this invisible Higgs decay and charged scalar contributions to h → γγ are generic features of
inert doublet schemes [29, 30] as well as scotogenic models [31, 32].
12

The ATLAS and CMS collaborations have studied this decay mode and their combined
analysis with 8 TeV data gives Rγγ exp
= 1.16+0.20
−0.18 [34]. For the 13 TeV Run-2, there is no
combined final data so far, and the available data is separated by production processes [35].
In our analysis we have used 13 TeV ATLAS result which gives the global signal strength
measurement of Rγγ exp
= 0.99+0.15
−0.14 [33]. In the left panel of Fig. 5 we show the invisible Higgs
branching ratio BR(h → η R η R ) as a function of the dark matter candidate mass mηR . In the
right panel we give the expected Rγγ values for the same random scan of parameters. One
sees that for low dark matter masses mηR < 60 GeV the invisible decay mode h → η R η R is
open and violates the LHC limit BR(h → Inv) ≤ 0.19 [28].
Likewise, the Rγγ measurement rules out the lower dark matter mass region. For inter-
mediate dark matter masses in the range 70 GeV ≤ mηR ≤ 100 GeV, there are acceptable
solutions with acceptable Rγγ ≈ 1. In the large mass region mηR > 550 GeV, the charged
Higgs η ± contribution to the diphoton decay mode h → γγ is negligible, so that Rγγ is
close to unity. From the above discussion, one can say that low mass dark matter with
mηR < 60 GeV is ruled out by LHC constraints. However, they can not completely rule out
the intermediate mass region 70 GeV ≤ mηR ≤ 100 GeV. Moreover, the heavy dark matter
mass region mηR > 550 GeV is completely allowed by LHC constraints.
Before concluding this section we should note that there are also constraints from LEP-I
and LEP-II experiments. The precise LEP-I measurements rule out SM-gauge bosons decays
to dark sector particles [30, 36]. This requires that

mηR + mηI , 2mη± > mZ , and mηR /ηI + mη± > mW (25)

Although there is no dedicated analysis of LEP-II data in the context of scotogenic dark
matter models, Ref. [37] has discussed LEP II limits for the case of the Inert Doublet Model,
∼ 80 GeV and mηI >
leading to the limits mηR > ∼ 100 GeV and a small ηR − ηI mass splitting.
Altogether, in view of the above, one can say that intermediate dark matter masses in the
range 80 GeV ≤ mηR ≤ 100 GeV are not inconsistent with collider constraints. Dark matter
heavier than 550 GeV is perfectly allowed. Dark matter masses in between 100 and 550
GeV could also be possible in the presence of another dark matter component.

6. ELECTROWEAK VACUUM STABILITY

The detailed analysis of the Higgs vacuum within the Standard Model has been carried
out in [38–43]. Taking into account the updated input top and Higgs boson mass values one
finds that the Standard Model Higgs quartic coupling λSM becomes negative at µ ' 1010
GeV. This would imply that the Higgs potential is unbounded from below and the Higgs
vacuum is unstable. A dedicated analysis shows that, actually, the Standard Model Higgs
vacuum is metastable with very long lifetime [43].
13

Here our aim is to determine the parameter region consistent both with dark matter
observations as well as vacuum stability. We first fix the parameters (quartic couplings and
mass of the dark matter candidate) which are consistent with the present day relic density
and the direct dark matter detection constraints.
In Fig. 6 we have shown the quartic coupling λ345 = λ3 + λ4 + λ5 as a function of dark
matter mass mηR . The color code is same as in Fig. 2. The lower mass region is already

FIG. 6: Coupling λ345 = λ3 + λ4 + λ5 as a function of the dark mass mηR . The color code
is same as in Fig. 2.

excluded by experiment. One sees that in the large mass regime above mηR > 550 GeV,
the allowed λ345 values that successfully explain the relic density while obeying the direct
detection limit cover a wide range. For relatively large couplings, the evolution of quartic
couplings can make them exceed the perturbativity limit even before the Planck scale. We
therefore choose relatively small λ3,4,5 values. This way λ345 is small enough to match the
required relic density and to satisfy the direct detection cross section bound for mηR > 550
GeV.

We now examine the effect of the new particles present in the scoto-seesaw model upon
the stability of the electroweak vacuum. As a “missing partner” (3,1) type-I seesaw cloned
with the simplest scotogenic sector, the scoto-seesaw contains the a “right-handed” neutrino
N , together with the dark particles f and η. Using SARAH [14] we have computed the
two-loop RGEs of the full theory for all the quartic scalar couplings, as well as Yukawa
couplings, as given in Appendix B.
We now summarise our results. To begin with, in the effective theory where the heavy
singlet fermions N and f are integrated out, we have two natural threshold scales ΛN ≈ MN
14

and Λf ≈ Mf . These masses are obtained from Eq. (4) and (5). Hence, in the RGEs
of the full theory, we can simply take the YN and Yf Yukawa parameters to be given as
θ (µ − MN ) YN , θ (µ − Mf ) Yf . Clearly they do not run in the effective theory.
Our aim is to study the effect of large as well as small Yukawa couplings on vacuum
stability. As discussed before, we take an adequately small but nonzero value for λ5 , as
required for having a reasonable direct detection cross section. For example, with Yukawa
couplings YN ∼ Yf ∼ O(1) and very large MN , Mf values, one sees from Eq. (18) that one
can easily reproduce the solar and atmospheric scale with mR η ∼ O(1 TeV).

We now illustrate in more detail the relevant parameter space of the scoto-seesaw model
which is consistent with vacuum stability as well as neutrino and dark matter phenomenol-
ogy. In order to do so we have chosen three sets of benchmarks, given as:

• BP1: MN ∼ 1014 GeV, Mf ∼ 1012 GeV, mηR ∼ 103 GeV, YN ∼ 0.45, Yf ∼ 0.45,

• BP2: MN ∼ 1014 GeV, Mf ∼ 105 GeV, mηR ∼ 103 GeV, YN ∼ 0.45, Yf ∼ 10−4 ,

• BP3: MN ∼ 106 GeV, Mf ∼ 106 GeV, mηR ∼ 103 GeV, YN ∼ 10−5 , Yf ∼ 10−4 .

In the upper left, upper right and bottom panel of Fig. 7, we have shown the results for
m2
benchmark points BP1, BP2 and BP3, respectively. We have taken λη = 0.1 and λ = 2vh2
at the electroweak scale. Recall that in order to have an absolutely stable vacuum, one
needs to satisfy the conditions given in Eqs. (7), (8) and (9) at all energy scales.

Assuming small λ5 we show in Fig. 7 the evolution of the remaining four quartic couplings
λ, λη , λA and λB . One sees that, with reasonable initial choices, all of the quartic couplings
can remain positive and perturbative all the way up to the Planck scale. Since the Yukawa
couplings have a negative effect in the RGE evolution of λ, the required value of the quartic
couplings λ3 and λ4 is correspondingly larger, as seen when going from the upper left to the
right panel and finally to the bottom panel in Fig. 7.

To sum up, one sees that the minimal scoto-seesaw model has improved stability prop-
erties compared to the type-I seesaw scenario, due to the new scalars needed to realize the
scotogenic “completion”. The model can explain both solar and atmospheric neutrino mass
scales as well as dark matter, upgrading the (3, 1) type-I seesaw mechanism, which can only
generate the atmospheric neutrino mass scale. Moreover, the minimal scoto-seesaw model
leads to a stable and perturbative electroweak vacuum all the way up to the Planck scale. It
can therefore be considered as a full consistent theory for neutrino masses and dark matter.
15

YN (ΛN )≈10-5, Yf (Λf )≈10-4, λ3=λ4=0.08 YN (ΛN )=0.45, Yf (Λf )≈10-4, λ3=λ4=0.09
0.5 0.5 YN

0.4 0.4 λB
λB
Couplings

Couplings
0.3 0.3 λA
λA
0.2 0.2

0.1 λη 0.1 λη
λ λ
0.0 0.0
103 106 109 1012 1015 1018 103 106 109 1012 1015 1018
μ [GeV] μ [GeV]
YN (ΛN )=0.45, Yf (Λf )=0.45, λ3=λ4=0.1
0.5 YN
λB Yf
0.4
λA
Couplings

0.3

0.2

0.1 λη
λ
0.0
103 106 109 1012 1015 1018
μ [GeV]

FIG. 7: The RGE-evolution of quartic couplings λ, λη , λA , λB defined in Eqs. (7),(8),(9). The Yukawas
YN , Yf are also shown. The values given in the boxes are the initial values at the respective threshold
scales. All couplings remain perturbative up to the Planck Scale. Note that we have fixed mηR = 1 TeV
and taken small λ5 = 10−3 . See text for more details.

7. HIGH ENERGY BEHAVIOR OF THE DARK PARITY

The conservation of the dark parity is a key feature of the scoto-seesaw model, ensuring
dark matter stability as well as the radiative origin of the solar mass scale. Without this
Z2 symmetry conservation, the LDP would no longer be stable, and also the solar neutrino
splitting would not be “calculable” from Eq. (18). It was first pointed out in Ref. [44] that
renormalization group evolution can alter the scalar potential at high energies, leading to
Z2 breaking. It is easy to understand the source of Z2 symmetry breaking from the one-
loop β function of the m2η parameter. This is given in Appendix B [45, 46]. One needs to
focus on the terms which contribute negatively to the evolution of m2η , given in Eq. (B17).
We see that with relatively large Yukawa coupling Yf (i.e. λ5  1) and Mf2 > m2η , the
term −|Mf |2 |Yf |2 dominates the running of m2η . This can quickly drive m2η towards negative
16

values and induce a minimum of the scalar potential with hηi 6= 0. Notice, however, that
there are terms in Eq. (B17) which can counter this negative effect. For example, terms
proportional to the quartic scalar couplings λ3 and λ4 may do so if their signs are properly
chosen. The contribution to the m2η evolution will be positive for λη > 0 and λ3,4 < 0.

Mf =103 GeV Mf =106 GeV


300 300

250 250 Yf =10-4


Yf =0.2

200 200
mη [GeV]

mη [GeV]

Yf =10

Yf=
150 Yf =0.5 Yf =0.3 150
Yf =3x10-4

5x1
Yf =0.4

-3
100 100

0
-4
50 50

0.0 0.0
103 106 109 1012 1015 1018 103 106 109 1012 1015 1018
μ [GeV] μ [GeV]

FIG. 8: RGE evolution of mη as a function of the energy scale µ. For both panels we start running at
mη = 250 GeV, taking YN = 0, λ3 = λ4 = 0.1 and λ5 = 10−3 . Different lines correspond to various Yukawa
coupling values, as indicated. Note that for Mf = 106 GeV (right panel) the Yukawas are much smaller.
Notice also that all curves terminate when m2η becomes negative (mη imaginary) indicating Z2 breakdown.

Fig. 8 shows the evolution of the scalar mass mη as a function of energy scale µ. The
results have been obtained for two values of Mf , Mf = 103 GeV (left panel) and 106
GeV (right panel). We have fixed m2η = 2502 GeV2 , λ3 = λ4 = 0.1, λ5 ≈ 0 and YN = 0, for
simplicity. The red-solid, blue-dot-dashed, orange-dashed and green-dotted lines in the left
(right) panel correspond to four values of the Yukawa coupling Yf , as indicated. As expected,
the Z2 breaking scale decreases for larger Mf due to the effect of the term −|Mf |2 |Yf |2 . In
other words, the larger the scale Mf , the smaller the allowed value of the Yukawa coupling
Yf in order to have the Z2 symmetry preserved all the way up to the Planck scale. From
Fig. 8, one sees that the allowed value of this Yukawa coupling is Yf ≤ 0.2 for Mf = 103
GeV, whereas for Mf = 106 GeV it is Yf ≤ 10−4 . Although the different quartic couplings
such as λ3 and λ4 may alter the details, this generic behavior remains. To sum up, we found
that in the scoto-seesaw model the dark parity can be preserved up to the Planck scale over
large portions of the parameter space.

8. SUMMARY AND DISCUSSION

We have examined the minimal combination of the the seesaw and scotogenic paradigms
for neutrino mass generation and dark matter able to “explain” the solar and atmospheric
17

oscillation wavelengths, in Eqs. 1 and 2. The model provides a simple picture where the
“atmospheric” mass scale arises from the tree-level “missing partner” seesaw, while the
“solar” scale is induced radiatively by the dark sector, see Fig. 1. We have derived the full
two-loop RGEs for the relevant parameters, such as the quartic Higgs self-coupling λ of
the Standard Model. The new scalars present in the scoto-seesaw mechanism improve the
stability properties of the electroweak vacuum, as seen in Fig. 7. We have also explored the
consistency of the underlying dark symmetry, as seen in Fig. 8.

Concerning phenomenology we have discussed scalar dark matter including the experi-
mental restrictions that follow from colliders, Fig. 5. By taking into account the relevant
annihiliation channels in Fig. 9 we identified viable parameter regions consistent with the
required dark matter relic abundance, Fig. 2. Direct dark matter detection by nucleon
recoil proceeds through the diagrams in Fig.3. The expected rates are given in Fig. 4 and
offer promising results for upcoming dark matter experiments. We found that the low dark
matter mass region mηR < ∼ 60 GeV is ruled out by LHC data, but the intermediate region
80 GeV < ∼ mηR <∼ 100 GeV is still allowed both by LHC and LEP data. The heavier mass
region mηR >∼ 550 GeV is free from collider constraints.

Our construction is very attractive from the point of view of neutrino physics. In contrast
to the original scotogenic model where 2 (or 3) species of either dark fermion or dark scalars
are needed to generate masses for 2 (or 3) neutrinos, our dark sector is truly minimal,
with only one dark fermion and one dark scalar. Therefore, the allowed parameter space
differs from the original scotogenic model, though these differences do not translate into
a phenomenological smoking-gun signature which can easily distinguish it from canonical
scotogenic model in Ref. [10].

Before closing we also note an important phenomenological implication of the minimal


scoto-seesaw model that can make it testable. Namely, it implies that one of the neutrinos
is massless, as can be readily seen from Eq. (16). This leads to a lower bound on the
neutrinoless double beta decay rates even for a normal-ordered neutrino mass spectrum [32,
47–49]. As shown in Fig.2 of Ref. [32], for an inverted mass spectrum the 0νββ lower
bound lies substantially higher than in the generic case for that ordering. As a result it falls
within the sensitivity of future experiments such as nEXO [50]. In the scoto-seesaw one has
that a positive 0νββ decay discovery would open bright prospects for the underpinning of
the value of the relevant elusive Majorana phase. As discussed in [4], the expected rates
for lepton flavour violation processes can also lie within reach of experiments, providing
additional signatures. In summary, the scoto-seesaw model is a theoretically consistent and
phenomenologicaly interesting “dark matter completion” of the type-I seesaw mechanism.
18

ACKNOWLEDGMENTS

This work is supported by the Spanish grant FPA2017-85216-P (AEI/FEDER, UE),


PROMETEO/2018/165 (Generalitat Valenciana). R.S. is supported by SERB, Government
of India grant SRG/2020/002303.

Appendix A: Scalar dark matter annihilation mechanisms

In the minimal scoto-seesaw model the relic abundance of the lightest dark particle η R is
determined by the following annihilation and coannihilation diagrams.

ηR h ηR ℓ− , q ηR ℓ− , q

h Z h
η R ηI ηR
h ℓ+ , q̄ ℓ+ , q̄
ηR ν ηR W+ ηR W+

W± h Z
±
η η R
ηI
ℓ± W− W−
ηR W− ηR W−

ηR W+ ηR Z
η± η±

ηR ηR
W − Z ηI W+ ηR W+
η R/I ν ηR h ηR ν

f ηR f

ηR ν̄ ηR h η± ℓ±

FIG. 9: Annihilation and coannihilation diagrams contributing to the relic abundance of ηR .


19

Appendix B: Renormalization group equations

The evolution of a given parameter c in the theory is described by the appropriate β


function, given by,
dc 1 (1) 1
≡ βc = βc + β (2) .
dt 16π 2 (16π 2 )2 c
(1) (2)
where βc are the one-loop renormalization group (RG) coefficients, while βc correspond
to the two-loop RG corrections.

1. Higgs quartic scalar self coupling

The scalar potential of the scoto-seesaw model is given in Eq. (6). The model contains
five quartic couplings λ, λη , λ3 , λ4 , λ5 . The one-loop and two-loop RG equations of the Higgs
quartic self-coupling λ are given by

(1) 27 4 9 9 9
βλ = + g1 + g12 g22 + g24 + 2λ23 + 2λ3 λ4 + λ24 + λ25 − g12 λ − 9g22 λ + 24λ2
200 20 8   5
2 † 4 † †
+ 12λyt + 4λTr YN YN − 6yt − 2Tr YN YN YN YN , (B1)

(2) 3537 6 1719 4 2 303 2 4 291 6 9 15 12


βλ = − g1 − g1 g2 − g1 g2 + g2 + g14 λ3 + g24 λ3 + g12 λ23 + 12g22 λ23
2000 400 80 16 10 2 5
9 3 15 12 6
− 8λ33 + g14 λ4 + g12 g22 λ4 + g24 λ4 + g12 λ3 λ4 + 12g22 λ3 λ4 − 12λ23 λ4 + g12 λ24
20 2 4 5 5
3 1953 117
+ 3g22 λ24 − 16λ3 λ24 − 6λ34 − g12 λ25 − 20λ3 λ25 − 22λ4 λ25 + g4λ + g2g2λ
5 200 1 20 1 2
51 108 2 2
− g24 λ − 20λ23 λ − 20λ3 λ4 λ − 12λ24 λ − 14λ25 λ + g λ + 108g22 λ2 − 312λ3
8 5 1
        171
− 4λ23 Tr Yf Yf† − 4λ3 λ4 Tr Yf Yf† − 2λ24 Tr Yf Yf† − 2λ25 Tr Yf Yf† − g4y2
100 1 t
63 9 17 45 9 4  
+ g12 g22 yt2 − g24 yt2 + g12 λyt2 + g22 λyt2 + 80g32 λyt2 − 144λ2 yt2 − g1 Tr YN YN†
10 4 2 2 100
3 2 2   3   3   15  
− g1 g2 Tr YN YN† − g24 Tr YN YN† + g12 λTr YN YN† + g22 λTr YN YN†
10 4 2 2
    8  
− 48λ2 Tr YN YN† − 3λTr Yf Yf† YN YN† − g12 yt4 − 32g32 yt4 − 3λyt4 − λTr YN YN† YN YN†
  5 
6 † † † † † †
+ 30yt + 2Tr Yf Yf YN YN YN YN + 10Tr YN YN YN YN YN YN . (B2)

(1) 27 4 9 4 9  
βλη = + g1 + g2 + 2λ23 + 2λ3 λ4 + λ24 + λ25 + g12 − 4λη + g22 − 9g22 λη + 24λ2η
200 8   20
† † †
+ 4λη Tr(Yf Yf ) − 2Tr( Yf Yf Yf Yf (B3)
20

(2) 3537 6 1719 4 2 303 2 4 291 6 9 15 12


βλη = − g1 − g1 g2 − g1 g2 + g2 + g14 λ3 + g24 λ3 + g12 λ23 + 12g22 λ23
2000 400 80 16 10 2 5
9 4 3 2 2 15 4 12 2 6
− 8λ3 + g1 λ4 + g1 g2 λ4 + g2 λ4 + g1 λ3 λ4 + 12g2 λ3 λ4 − 12λ23 λ4 + g12 λ24
3 2
20 2 4 5 5
2 2 2 3 3 2 2 2 2 1953 4 117 2 2
+ 3g2 λ4 − 16λ3 λ4 − 6λ4 − g1 λ5 − 20λ3 λ5 − 22λ4 λ5 + g λη + g g λη
5 200 1 20 1 2
51 108 2 2
− g24 λη − 20λ23 λη − 20λ3 λ4 λη − 12λ24 λη − 14λ25 λη + g1 λη + 108g22 λ2η − 312λ3η
8     5   
− λη Tr Yf Yf Yf Yf + 10Tr( Yf Yf Yf Yf Yf Yf − 3λη Tr Yf YN† YN Yf† − 4λ23 Tr YN YN†
† † † † †

         
† 2 † 2 † † † †
− 4λ3 λ4 Tr YN YN − 2λ4 Tr YN YN − 2λ5 Tr YN YN + Tr Yf Yf 2Tr Yf YN YN Yf
3  2    
− 10g1 − 5λη + g22 + 25 − 10g22 λη + 64λ2η + g24 + 3g14 − 12λ23 yt2
100
− 12λ3 λ4 yt2 − 6λ24 yt2 − 6λ25 yt2 (B4)

(1) 9 2 2 9 2 2 2 2



βλ4 = + g1 g2 − g1 λ4 − 9g2 λ4 + 8λ3 λ4 + 4λ4 + 8λ5 + 4λ4 λη + 4λ4 λ + 2λ4 Tr Yf Yf
5 5   

− 4Tr Yf YN YN Yf† + 2λ4 Tr YN YN† + 6λ4 yt2 (B5)

(2) 657 4 2 42 2 4 6 2 2 1413 4 153 2 2 231 4 12


βλ4 = − g1 g2 − g1 g2 + g1 g2 λ3 + g1 λ4 + g1 g2 λ4 − g2 λ4 + g12 λ3 λ4
50 5 5 200 20 8 5
24 48
+ 36g22 λ3 λ4 − 28λ23 λ4 + g12 λ24 + 18g22 λ24 − 28λ3 λ24 + g12 λ25 + 54g22 λ25 − 48λ3 λ25
5 5
24
− 26λ4 λ25 + 6g12 g22 λη + g12 λ4 λη − 80λ3 λ4 λη − 40λ24 λη − 48λ25 λη − 28λ4 λ2η + 6g12 g22 λ
5
24 2 9    
+ g1 λ4 λ − 80λ3 λ4 λ − 40λ24 λ − 48λ25 λ − 28λ4 λ2 − λ4 Tr Yf Yf† Yf Yf† + Tr YN YN†
5 2
 3 3 15  9  
− g12 g22 + g12 λ4 + g22 λ4 − 8λ3 λ4 − 4λ24 − 8λ25 − 8λ4 λ − λ4 Tr YN YN† YN YN†
5 4 4 2
      3 3
+ Tr Yf YN† YN Yf† − 3λ4 + 8λ3 + 8Tr YN YN† + Tr Yf Yf† − g12 g22 + g12 λ4
5 4
15 2    63 17
+ g2 λ4 − 8λ3 λ4 − 4λ24 − 8λ25 − 8λ4 λη + 8Tr Yf YN† YN Yf† + yt2 g12 g22 + g12 λ4
4 5 4
45 2 27 
+ g2 λ4 + 40g32 λ4 − 24λ3 λ4 − 12λ24 − 24λ25 − 24λ4 λ − λ4 yt2 (B6)
4 2

(1) 27 4 9 9 9
βλ3 = + g1 − g12 g22 + g24 − g12 λ3 − 9g22 λ3 + 4λ23 + 2λ24 + 2λ25 + 12λ3 λη
100 10 4 5    
+ 4λ4 λη + 12λ3 λ + 4λ4 λ + 2λ3 Tr Yf Yf† + 2λ3 Tr YN YN† + 6λ3 yt2 (B7)
21

(2) 3537 6 909 4 2 33 2 4 291 6 1773 4 33 111 4 6


βλ3 = − g1 + g1 g2 + g1 g2 + g2 + g1 λ3 + g12 g22 λ3 − g2 λ3 + g12 λ23
1000 200 40 8 200 20 8 5
9 9 15 6
+ 6g22 λ23 − 12λ33 + g14 λ4 − g12 g22 λ4 + g24 λ4 − 12g22 λ3 λ4 − 4λ23 λ4 − g12 λ24
10 5 2 5
12 27
+ 6g22 λ24 − 16λ3 λ24 − 12λ34 + g12 λ25 − 18λ3 λ25 − 44λ4 λ25 + g14 λη − 3g12 g22 λη
5 10
45 4 72 2 24
+ g2 λη + g1 λ3 λη + 72g22 λ3 λη − 72λ23 λη + g12 λ4 λη + 36g22 λ4 λη − 32λ3 λ4 λη
2 5 5
27 45 72
− 28λ24 λη − 36λ25 λη − 60λ3 λ2η − 16λ4 λ2η + g14 λ − 3g12 g22 λ + g24 λ + g12 λ3 λ
10 2 5
24
+ 72g22 λ3 λ − 72λ23 λ + g12 λ4 λ + 36g22 λ4 λ − 32λ3 λ4 λ − 28λ24 λ − 36λ25 λ − 60λ3 λ2
5
9   9 4   3   3  
− 16λ4 λ2 − λ3 Tr Yf Yf† Yf Yf† − g1 Tr YN YN† + g12 g22 Tr YN YN† − g24 Tr YN YN†
2 100 10 4
3 2   15        
+ g1 λ3 Tr YN YN† + g22 λ3 Tr YN YN† − 4λ23 Tr YN YN† − 2λ24 Tr YN YN† − 2λ25 Tr YN YN†
4 4
    9     9 4
− 24λ3 λTr YN YN† − 8λ4 λTr YN YN† − λ3 Tr YN YN† YN YN† + Tr Yf Yf† − g
2 100 1
3 3 3 15  
+ g12 g22 − g24 + g12 λ3 + g22 λ3 − 4λ23 − 2λ24 − 2λ25 − 24λ3 λη − 8λ4 λη + 4Tr Yf YN† YN Yf†
10 4 4 4
        171 63 9
+ Tr Yf YN† Tr YN Yf† − 3λ3 + 4Tr YN YN† + 8λ4 + yt2 − g14 − g12 g22 − g24
100 10 4
17 2 45 2 27 
+ g1 λ3 + g2 λ3 + 40g32 λ3 − 12λ23 − 6λ24 − 6λ25 − 72λ3 λ − 24λ4 λ − λ3 yt2 (B8)
4 4 2

9    
βλ5 = − g12 λ5 − 9g22 λ5 + 8λ3 λ5 + 12λ4 λ5 + 4λ5 λη + 4λ5 λ + 2λ5 Tr Yf Yf† + 2λ5 Tr YN YN†
(1)
5
+ 6λ5 yt2 (B9)

(2) 1413 4 57 231 4 48 72


βλ5 = + g1 λ5 + g12 g22 λ5 − g2 λ5 + g12 λ3 λ5 + 36g22 λ3 λ5 − 28λ23 λ5 + g12 λ4 λ5
200 20 8 5 5
12
+ 72g22 λ4 λ5 − 76λ3 λ4 λ5 − 32λ24 λ5 + 6λ35 − g12 λ5 λη − 80λ3 λ5 λη − 88λ4 λ5 λη − 28λ5 λ2η
5
12 2 1      
− g1 λ5 λ − 80λ3 λ5 λ − 88λ4 λ5 λ − 28λ5 λ2 + λ5 15g22 − 16 2λ3 + 2λη + 3λ4 + 3g12 Tr Yf Yf†
5 4
1     3   15  
− λ5 Tr Yf Yf† Yf Yf† − 3λ5 Tr Yf YN† YN Yf† + g12 λ5 Tr YN YN† + g22 λ5 Tr YN YN†
2 4 4
      1  
− 8λ3 λ5 Tr YN YN† − 12λ4 λ5 Tr YN YN† − 8λ5 λTr YN YN† − λ5 Tr YN YN† YN YN†
2
 17 45 3 
+ yt2 g12 λ5 + g22 λ5 + 40g32 λ5 − 24λ3 λ5 − 36λ4 λ5 − 24λ5 λ − λ5 yt2 (B10)
4 4 2
22

2. Yukawa Couplings

The one-loop and two-loop RG equations for the Yukawa couplings Yf , YN and yt are
given by

(1) 1       
βYf = 10 3Yf Yf† Yf + YN YN† Yf + Yf 20Tr Yf Yf† − 9 5g22 + g12 , (B11)
20

(2) 1 2
βYf = + 33g1 YN YN† Yf + 165g22 YN YN† Yf − 160λ3 YN YN† Yf − 320λ4 YN YN† Yf
80
+ 120Yf Yf† Yf Yf† Yf − 20Yf Yf† YN YN† Yf − 20YN YN† YN YN† Yf − 180YN YN† Yf yt2
     117 27
† 2 † 2
+ 3Yf Yf Yf 225g2 − 320λη − 60Tr Yf Yf + 93g1 + Yf g14 − g12 g22
200 20
21 4 2 2 3 2 2 3 2 2
 


− g2 + λ3 + λ3 λ4 + λ4 + λ5 + 6λη + 5g2 + g1 Tr Yf Yf
4 2 8
9   3  
− Tr Yf Yf† Yf Yf† − Tr Yf Yf† YN YN† , (B12)
4 4

(1) 1 † †
 
2 9 2 9 2 


β YN = + 3YN YN YN + Yf Yf YN + YN 3yt − g1 − g2 + Tr YN YN , (B13)
2 20 4

(2) 1
βYN = + 279g12 YN YN† YN + 675g22 YN YN† YN − 960λYN YN† YN − 20Yf Yf† Yf Yf† YN
80 
† † † † †
− 20YN YN Yf Yf YN + 120YN YN YN YN YN + Yf Yf YN − 160λ3 + 165g22 − 320λ4
   
+ 33g12 − 60Tr Yf Yf† − 540YN YN† YN yt2 − 180YN YN† YN Tr YN YN†
 117 27 21 3 17
+ YN g14 − g12 g22 − g24 + λ23 + λ3 λ4 + λ24 + λ25 + 6λ2 + g12 yt2
200 20 4 2 8
45 2 2 2 2 3 2


 15
2


 3 
† †

+ g2 yt + 20g3 yt + g1 Tr YN YN + g2 Tr YN YN − Tr Yf Yf YN YN
8 8 8 4
27 4 9  † †

− yt − Tr YN YN YN YN , (B14)
4 4

3 3  17 2 9 2 


βy(1) = y + y t 3y 2
− 8g 2
− g − g + Tr Y N Y , (B15)
t
2 t t 3
20 1 4 2 N

1  

 
βy(2)
t
= + 120y 5
t + y 3
t 1280g 2
3 − 180Tr Y Y
N N + 223g1
2
− 540yt
2
+ 675g 2
2 − 960λ
80
 1267 9 21 19 3
+ yt g14 − g12 g22 − g24 + g12 g32 + 9g22 g32 − 108g34 + λ23 + λ3 λ4 + λ24 + λ25
600 20 4 15 2
17 45 3   15  
+ 6λ2 + g12 yt2 + g22 yt2 + 20g32 yt2 + g12 Tr YN YN† + g22 Tr YN YN†
8 8 8 8
3   27 9  
− Tr Yf Yf† YN YN† − yt4 − Tr YN YN† YN YN† (B16)
4 4 4
23

3. Scalar Mass term

The evolution of the scalar mass-squared term m2η is dictated by the β function

(1)   † 
βm2η = 12λη mη + 2 −2|Mf | + mη Tr Yf Yf − 2(λ4 + 2λ3 )µ2H
2 2 2

 
9 2 9 2
− g + g m2η (B17)
10 1 2 2

[1] A. B. McDonald, “Nobel Lecture: The Sudbury Neutrino Observatory: Observation of flavor
change for solar neutrinos,” Rev. Mod. Phys. 88 no. 3, (2016) 030502.
[2] T. Kajita, “Nobel Lecture: Discovery of atmospheric neutrino oscillations,” Rev. Mod. Phys.
88 no. 3, (2016) 030501.
[3] G. Bertone, D. Hooper, and J. Silk, “Particle dark matter: Evidence, candidates and
constraints,” Phys.Rept. 405 (2005) 279–390.
[4] N. Rojas, R. Srivastava, and J. W. F. Valle, “Simplest Scoto-Seesaw Mechanism,” Phys.
Lett. B789 (2019) 132–136, arXiv:1807.11447 [hep-ph].
[5] P. F. de Salas et al., “2020 global reassessment of the neutrino oscillation picture,” JHEP 02
(2021) 071, arXiv:2006.11237 [hep-ph].
[6] P. F. De Salas et al., “Chi2 profiles from Valencia neutrino global fit.”
http://globalfit.astroparticles.es/, 2021.
https://doi.org/10.5281/zenodo.4593330.
[7] D. Barreiros, F. Joaquim, R. Srivastava, and J. W. F. Valle, “Minimal scoto-seesaw
mechanism with spontaneous CP violation,” arXiv:2012.05189 [hep-ph].
[8] J. Schechter and J. W. F. Valle, “Neutrino Masses in SU(2) x U(1) Theories,” Phys. Rev.
D22 (1980) 2227.
[9] J. Schechter and J. W. F. Valle, “Neutrino Decay and Spontaneous Violation of Lepton
Number,” Phys. Rev. D25 (1982) 774.
[10] E. Ma, “Verifiable radiative seesaw mechanism of neutrino mass and dark matter,”
Phys.Rev. D73 (2006) 077301, arXiv:hep-ph/0601225 [hep-ph].
[11] M. Hirsch, M. Diaz, W. Porod, J. Romao, and J. W. F. Valle, “Neutrino masses and mixings
from supersymmetry with bilinear R parity violation: A Theory for solar and atmospheric
neutrino oscillations,” Phys.Rev. D62 (2000) 113008.
[12] M. Diaz et al., “Solar neutrino masses and mixing from bilinear R parity broken
supersymmetry: Analytical versus numerical results,” Phys.Rev. D68 (2003) 013009.
[13] M. Hirsch and J. W. F. Valle, “Supersymmetric origin of neutrino mass,” New J.Phys. 6
(2004) 76.
24

[14] F. Staub, “Exploring new models in all detail with SARAH,” Adv. High Energy Phys. 2015
(2015) 840780, arXiv:1503.04200 [hep-ph].
[15] G. Bélanger, F. Boudjema, A. Goudelis, A. Pukhov, and B. Zaldivar, “micrOMEGAs5.0 :
Freeze-in,” Comput. Phys. Commun. 231 (2018) 173–186, arXiv:1801.03509 [hep-ph].
[16] Planck Collaboration, N. Aghanim et al., “Planck 2018 results. VI. Cosmological
parameters,” arXiv:1807.06209 [astro-ph.CO].
[17] XENON Collaboration, E. Aprile et al., “Dark Matter Search Results from a One Ton-Year
Exposure of XENON1T,” Phys. Rev. Lett. 121 no. 11, (2018) 111302, arXiv:1805.12562
[astro-ph.CO].
[18] PandaX Collaboration, H. Zhang et al., “Dark matter direct search sensitivity of the
PandaX-4T experiment,” Sci. China Phys. Mech. Astron. 62 no. 3, (2019) 31011,
arXiv:1806.02229 [physics.ins-det].
[19] LUX-ZEPLIN Collaboration, D. S. Akerib et al., “Projected WIMP sensitivity of the
LUX-ZEPLIN dark matter experiment,” Phys. Rev. D 101 no. 5, (2020) 052002,
arXiv:1802.06039 [astro-ph.IM].
[20] XENON Collaboration, E. Aprile et al., “Projected WIMP sensitivity of the XENONnT
dark matter experiment,” JCAP 11 (2020) 031, arXiv:2007.08796 [physics.ins-det].
[21] GADMC Collaboration, C. Galbiati et al., “Future Dark Matter Searches with
Low-Radioactivity Argon,” Input to the European Particle Physics Strategy Update
2018-2020 (2018) . https://indico.cern.ch/event/765096/contributions/3295671/
attachments/1785196/2906164/DarkSide-Argo_ESPP_Dec_17_2017.pdf.
[22] DARWIN Collaboration, J. Aalbers et al., “DARWIN: towards the ultimate dark matter
detector,” JCAP 1611 (2016) 017, arXiv:1606.07001 [astro-ph.IM].
[23] J. Billard et al., “Direct Detection of Dark Matter – APPEC Committee Report,”
arXiv:2104.07634 [hep-ex].
[24] J. Billard, L. Strigari, and E. Figueroa-Feliciano, “Implication of neutrino backgrounds on
the reach of next generation dark matter direct detection experiments,” Phys. Rev. D 89
no. 2, (2014) 023524, arXiv:1307.5458 [hep-ph].
[25] LUX Collaboration, D. Akerib et al., “Results from a search for dark matter in the complete
LUX exposure,” Phys.Rev.Lett. 118 (2017) 021303, arXiv:1608.07648 [astro-ph.CO].
[26] PandaX-II Collaboration, A. Tan et al., “Dark Matter Results from First 98.7 Days of
Data from the PandaX-II Experiment,” Phys. Rev. Lett. 117 no. 12, (2016) 121303,
arXiv:1607.07400 [hep-ex].
[27] A. Pierce and J. Thaler, “Natural Dark Matter from an Unnatural Higgs Boson and New
Colored Particles at the TeV Scale,” JHEP 08 (2007) 026, arXiv:hep-ph/0703056.
[28] CMS Collaboration, A. M. Sirunyan et al., “Search for invisible decays of a Higgs boson

produced through vector boson fusion in proton-proton collisions at s = 13 TeV,” Phys.
25

Lett. B 793 (2019) 520–551, arXiv:1809.05937 [hep-ex].


[29] R. Barbieri, L. J. Hall, and V. S. Rychkov, “Improved naturalness with a heavy Higgs: An
Alternative road to LHC physics,” Phys. Rev. D 74 (2006) 015007, arXiv:hep-ph/0603188.
[30] Q.-H. Cao, E. Ma, and G. Rajasekaran, “Observing the Dark Scalar Doublet and its Impact
on the Standard-Model Higgs Boson at Colliders,” Phys. Rev. D 76 (2007) 095011,
arXiv:0708.2939 [hep-ph].
[31] C. Bonilla, E. Ma, E. Peinado, and J. W. F. Valle, “Two-loop Dirac neutrino mass and
WIMP dark matter,” Phys.Lett. B762 (2016) 214–218, arXiv:1607.03931 [hep-ph].
[32] I. M. Ávila, V. De Romeri, L. Duarte, and J. W. F. Valle, “Phenomenology of scotogenic
scalar dark matter,” Eur. Phys. J. C 80 no. 10, (2020) 908, arXiv:1910.08422 [hep-ph].
[33] ATLAS Collaboration, M. Aaboud et al., “Measurements of Higgs boson properties in the

diphoton decay channel with 36 fb−1 of pp collision data at s = 13 TeV with the ATLAS
detector,” Phys. Rev. D 98 (2018) 052005, arXiv:1802.04146 [hep-ex].
[34] ATLAS, CMS Collaboration, G. Aad et al., “Measurements of the Higgs boson production
and decay rates and constraints on its couplings from a combined ATLAS and CMS analysis

of the LHC pp collision data at s = 7 and 8 TeV,” JHEP 08 (2016) 045,
arXiv:1606.02266 [hep-ex].
[35] ATLAS Collaboration, G. Aad et al., “Combined measurements of Higgs boson production

and decay using up to 80 fb−1 of proton-proton collision data at s = 13 TeV collected with
the ATLAS experiment,” Phys. Rev. D 101 no. 1, (2020) 012002, arXiv:1909.02845
[hep-ex].
[36] M. Gustafsson, E. Lundstrom, L. Bergstrom, and J. Edsjo, “Significant Gamma Lines from
Inert Higgs Dark Matter,” Phys. Rev. Lett. 99 (2007) 041301, arXiv:astro-ph/0703512.
[37] E. Lundstrom, M. Gustafsson, and J. Edsjo, “The Inert Doublet Model and LEP II Limits,”
Phys. Rev. D 79 (2009) 035013, arXiv:0810.3924 [hep-ph].
[38] G. Isidori, G. Ridolfi, and A. Strumia, “On the metastability of the standard model
vacuum,” Nucl.Phys. B609 (2001) 387–409.
[39] J. Elias-Miro, J. R. Espinosa, G. F. Giudice, G. Isidori, A. Riotto, and A. Strumia, “Higgs
mass implications on the stability of the electroweak vacuum,” Phys.Lett. B709 (2012)
222–228, arXiv:1112.3022 [hep-ph].
[40] F. Bezrukov, M. Y. Kalmykov, B. A. Kniehl, and M. Shaposhnikov, “Higgs Boson Mass and
New Physics,” vol. 1210, p. 140. 2012. arXiv:1205.2893 [hep-ph].
[41] G. Degrassi et al., “Higgs mass and vacuum stability in the Standard Model at NNLO,”
JHEP 1208 (2012) 098, arXiv:1205.6497 [hep-ph].
[42] I. Masina, “Higgs boson and top quark masses as tests of electroweak vacuum stability,”
Phys.Rev. D87 (2013) 053001, arXiv:1209.0393 [hep-ph].
26

[43] D. Buttazzo, G. Degrassi, P. P. Giardino, G. F. Giudice, F. Sala, A. Salvio, and A. Strumia,


“Investigating the near-criticality of the Higgs boson,” JHEP 12 (2013) 089,
arXiv:1307.3536 [hep-ph].
[44] A. Merle and M. Platscher, “Parity Problem of the Scotogenic Neutrino Model,” Phys. Rev.
D 92 no. 9, (2015) 095002, arXiv:1502.03098 [hep-ph].
[45] A. Merle and M. Platscher, “Running of radiative neutrino masses: the scotogenic model —
revisited,” JHEP 11 (2015) 148, arXiv:1507.06314 [hep-ph].
[46] M. Lindner, M. Platscher, C. E. Yaguna, and A. Merle, “Fermionic WIMPs and vacuum
stability in the scotogenic model,” Phys. Rev. D 94 no. 11, (2016) 115027,
arXiv:1608.00577 [hep-ph].
[47] M. Reig, D. Restrepo, J. W. F. Valle, and O. Zapata, “Bound-state dark matter with
Majorana neutrinos,” Phys.Lett. B790 (2019) 303–307, arXiv:1806.09977 [hep-ph].
[48] D. Barreiros, R. Felipe, and F. Joaquim, “Combining texture zeros with a remnant CP
symmetry in the minimal type-I seesaw,” JHEP 1901 (2019) 223, arXiv:1810.05454
[hep-ph].
[49] J. Leite et al., “A theory for scotogenic dark matter stabilised by residual gauge symmetry,”
arXiv:1909.06386 [hep-ph].
[50] M. J. Jewell, Search for neutrinoless double beta decay with EXO-200 and nEXO. PhD
thesis, Stanford U., 2020. https://stacks.stanford.edu/file/druid:
kf548hj0975/jewell_thesis_final_5_1_2020-augmented.pdf.

You might also like