You are on page 1of 10

Journal of Visualization manuscript No.

(will be inserted by the editor)

Structural Reynolds analogy in laminarescent boundary layers via DNS

G. Araya1,∗ · G. Torres 1

Abstract Visualization of the thermal field in highly-accelerated spatially developing turbulent boundary layers is
carried out. Direct Numerical Simulation (DNS) with high spatial/temporal resolution is performed in sink flow-type
boundary layer by prescribing a passive scalar with a Prandtl number of 0.71. The range of the momentum thickness
Reynolds number is approximately 320-432. The very strong Favorable Pressure Gradient (FPG) is imposed by a top
converging shearless surface, which produces an approximately constant acceleration parameter of K = 4.0 × 10−6 . A
precursor zone is prescribed upstream of the FPG region in order to generate accurate turbulent inflow information
by means of the methodology proposed by [7]. While evident “signatures” of the very strong FPG have been identified
in the velocity field, those “signatures” are much less evident in the temperature field causing a breakdown of the
Reynolds analogy provoked by the streamwise pressure gradient (or source of dissimilarity between the momentum
and thermal transport). A slow decrease of the thermal boundary layer thickness, δT , is observed in the FPG zone.
Additionally, local maxima of absolute intensities of the thermal fluctuations t0RM S exhibit nearly constant values
(aprox. 14% of T∞ ) along the composite domain. Conversely, absolute intensities of the streamwise (u0RM S ) show
0 0
increases on their local maxima, with mild decreases in the transversal components of the velocity (vRM S and wRM S ).
0 0
Furthermore, profiles of cross-correlation < u t > at different streamwise stations depict a good collapsing level up to
4% of δ in the wall-normal direction, with peak values displacing farther from the wall in the FPG zone.
Keywords DNS · Reynolds analogy · quasi-laminarization · passive scalar
PACS 47.27.nb · 47.27.N-

1 Introduction
Turbulence is mainly characterized by randomness, chaos and a disparate range of turbulent scales and mixing. Despite
its chaotic behavior, investigations performed during the last six decades has conclusively demonstrated the presence
of well organized turbulent motions in wall bounded flows, so called Coherent Structures (CS). These structures can
be considered the building-blocks of turbulent boundary layers and are responsible for carrying most of the turbulent
kinetic energy as well as the Reynolds stresses of the flow, Guala et al. [14]. Furthermore, a passive scalar is defined as
a diffusive contaminant that exists in such a low concentration in a turbulent flow that has no effect on the dynamics of
the fluid motion, Warhaft [31]. However, a low concentration of passive scalar is sufficient to provoke a significant im-
pact on energy expenditures, heat transfer, air pollution, and design of chemical processes. The transport phenomenon
in real-situation flows usually occurs under complicated external conditions, such as streamwise pressure gradients
and spatially-developing boundary layers. In this sense, Direct Numerical Simulation (DNS) with millions of “flow
sensors” and high temporal resolution may shed important light on the unknown aspects of the transport phenomena
in accelerated boundary layers. Coherent structures in such a complex environment and their interactions (turbulent
events) are better identified and visualized by DNS (Adrian and Liu [3], Mito and Kasagi [19]). Several studies on
passive scalar transport in turbulent channel flows have been performed. A numerical analysis of the effects of steady
and unsteady blowing/suction by means of 2D spanwise slots on the passive scalar field in a turbulent channel was
introduced by Araya et al. [8]. Authors found a characteristic forcing frequency of f + = 0.044 in wall units at which
the small turbulent scales were better excited by unsteady blowing/suction, while large scale motions were enhanced
by steady perturbations. Moreover, the coherent component of fluctuations exhibited large values very close to the slot
and in the near wall region; although they exhibited a significant attenuation downstream. Antonia et al. [4] studied
the spectral analogy between the fluctuating velocity vector and thermal fluctuations for various Reynolds numbers,
a Prandtl number of 0.71 and iso-flux wall conditions by means of DNS. While they found a reasonable similarity
between the spectra of kinetic energy and scalar variance in both inner and outer regions, the spectra of enstrophy and
scalar dissipation rate did not show a similar trend. Iztok [30] concluded that large-scale motions played a key role in
the temperature fluctuations at low Prandtl number. Recently, Pirozzoli et al. [26] carried out DNS of passive scalars
at high friction Reynolds numbers (up to Reτ = huτ /ν = 4088, where h is the channel half-height, uτ is the friction
velocity and ν is the kinematic fluid viscosity). It was concluded that the scalar dissipation was significantly larger
than the streamwise velocity dissipation, attributed to the presence of more energetic small scalar-bearing eddies, as
1 High Performance Computing and Visualization Laboratory

Department of Mechanical Engineering, University of Puerto Rico, Mayaguez, PR 00681, US


∗ Corresponding author: araya@mailaps.org
2 Structural Reynolds analogy

confirmed by analysis of the energy spectra and instantaneous flow visualizations.


One of the main challenges on the numerical simulations of spatially-developing boundary layers is the prescription
of realistic time-dependent inflow information. An extensive literature review about different inflow generation tech-
niques can be found in [33]. To the best of our knowledge, there are only few publications on DNS of turbulent thermal
spatially-evolving boundary layers in incompressible flows. For instance, Bell and Ferziger [11] performed DNS in
iso-scalar walls for an extent of momentum thickness Reynolds number from 300 to 700 and values of the Prandtl
number equal to 0.1, 0.71, and 2.0. Later, Kong et al. [16] extended the rescaling-recycling method by [18] to generate
time-dependant inflow thermal information in direct simulations of turbulent boundary layers in Zero Pressure Gra-
dient (ZPG) flows. Li et al. [17] carried out DNS of a spatially developing turbulent boundary layer over a flat plate
with the evolution of several passive scalars (Prandtl numbers = 0.2, 0.71, and 2) under both isoscalar and isoflux
wall boundary conditions. Wu & Moin [32] have reported DNS of an incompressible ZPG thermal boundary layer
with isothermal wall conditions. The range for the momentum thickness Reynolds numbers, Reθ , was 80 - 1950 and
the molecular Prandtl number Pr = 1. More recently, Zhao et al. [34] conducted DNS of an isothermal-flat plate for
momentum thickness Reynolds numbers of 80 to 1150 and a Prandtl number of 0.71. Furthermore, the DNS studies of
turbulent thermal spatially-evolving boundary layers mentioned so far have only involved ZPG flows. Araya & Castillo
([6], [13]) have investigated the effects of pressure gradients on thermal spatially-developing boundary layers by means
of DNS . They detected a clear “plateau” between the inner and outer peaks for streamwise velocity fluctuations in
strong Adverse Pressure Gradient (APG) flows. On the other hand, the outer peak was not very evident on thermal
fluctuations, which was attributed to the lack of a freestream thermal gradient. Furthermore, the high-speed streaks in
the buffer layer became significantly shorter and wider in a strong APG than in the ZPG case. In terms of FPG flows,
the acceleration parameter K is generally used for characterization purposes. An interesting problem in fluid dynamics
is when an initially turbulent boundary layer is subject to a sufficiently high FPG to induce “soft” relaminarization or
“quasi-laminarization”, in which “there may be appreciable residual turbulence” (Narasimha [21]). For K > 3.0×10−6 ,
relaminarization in turbulent flows is expected. However, Spalart [28] performed DNS of quasi-homogeneous sink flows
and obtained laminarization at K = 3.0 × 10−6 . Kline et al. [15] measured bursting in a boundary layer subjected to
strong FPG and proposed a critical acceleration parameter, Kc , of 3.5 × 10−6 at which bursting ceased. Significant
research effort has been invested during the last few decades to propose and evaluate a reliable flow parameter in
order to accurately describe the onset of relaminarization [15, 22, 29, 28, 25]. To our knowledge, the majority of previous
studies has focused on the influence of quasi-laminarization on the velocity field, whereas investigation of the effects
on the thermal transport is nearly non-existent.
Therefore, we seek to evaluate the structural Reynolds analogy between the hydrodynamic and thermal field in tur-
bulent boundary layers at an early stage of quasi-laminarization (i.e., laminarescent stage) induced by a very strong
FPG. Extensive information supplied by DNS is employed to carry out scientific visualization on turbulent structures
and events as well as to analyze cross-correlations of the velocity-temperature field (turbulent heat fluxes).

2 Numerical details
In recent years, the discipline of fluid dynamics has been reliant
to high-performance computational simulations as a means of pre-
dicting flow behavior and better understanding of the dynamics
of coherent motions. DNS is a numerical tool that resolves all
turbulence scales; thus, it aims to provide high spatial/temporal
thermal-fluid data within a computational domain. Furthermore,
turbulent boundary layers that evolve along the flow direction are
ubiquitous. Computationally speaking, these types of boundary
layers (i.e., spatially-developing boundary layers) pose an enor-
mous challenge, due to the need for accurate and time dependent
inflow turbulence information. Moreover, accounting for the effects
of streamwise pressure gradient adds significant complexity to the
problem. We are making use of the Dynamic Multi-scale Approach
(Araya et al. [7]), a method for prescribing realistic turbulent ve-
locity inflow boundary conditions. It is based on the rescaling-
recycling method proposed by Lund et al. [18]. The seminal idea
of the rescaling-recycling method is to prescribe time-dependent
turbulent information at the inlet plane based on the transformed Fig. 1: Schematic of the thermal boundary layer.
flow solution downstream, from a plane called recycle, by using
scaling laws. Furthermore, in our proposed new approach there is no need to use empirical correlations to compute
inlet parameters, as in the methodology introduced by [18]. To calculate the inlet friction velocity (uτ ) and friction
temperature (θτ ), an additional plane is involved, the so called test plane located between the inlet and recycle sta-
tions (Fig. 1). An extension of the Dynamic Multi-scale Approach to thermal boundary layers has been proposed
in [6] and [13]. The computational domain consists of a ZPG region (or precursor zone for inflow turbulent infor-
Structural Reynolds analogy in laminarescent boundary layers via DNS 3

(a) (b)

Fig. 2: Kolmogorov and Obukhoff-Corrsin length scales at (a) x/δinlet = 10 and (b) x/δinlet = 55.

mation generation) of 20δinlet -length followed by a FPG region of 40δinlet -length, where δinlet is the measured 99%
boundary layer thickness at the inlet. Favorable pressure gradient is prescribed by a top converging surface (sink flow)
with an approximately constant acceleration parameter of K = 4.0 × 10−6 . The acceleration parameter is defined as
K = Uν2 dU ∞
, where U∞ is the freestream velocity and ν is the fluid kinematic viscosity. The inlet, test and recycle
∞ dx
planes are located in the ZPG region, immediately preceding the FPG region. At the wall, the classical no-slip con-
dition is imposed for all velocity components. An isothermal condition is prescribed for the temperature field at the
wall, which is regarded as a passive scalar. The lateral boundary conditions are handled via periodicity. Pressure is
weakly prescribed at the outlet plane, whereas top surfaces are shearless. Dimensions of the composite computational
domain (Lx , Ly and Lz ) are 60δinlet , 4.3δinlet and 4.3δinlet along the streamwise (x), wall-normal (y) and spanwise
(z) directions, respectively. The mesh configuration is 600 × 80 × 80, which represents the numbers of points along x,
+ + +
y and z directions, respectively. The mesh resolution in wall units
p is ∆x = ∆x uτ /ν = 15, ∆y min = 0.2, ∆y max =
+
13 and ∆z = 8. Here, the friction velocity is defined as uτ = τw /ρ, where τw is the wall shear stress and ρ is the
fluid density. The Courant-Friedrichs-Levy (CFL) parameter is fixed at 0.24 during the simulation and the time step
is ∆t + ≈ 0.19. More numerical details can be found in Araya et al. [5]. They situated this flow in the laminarescent
region, an earlier stage of quasi-laminarization according to Narasimha & Sreenivasan [22].

3 Results and discussion

Suitability of the mesh resolution is demonstrated by means of the computation of the Kolmogorov and the Obukhoff-
Corrsin length scales, for the velocity and thermal field, respectively. Furthermore, the Kolmogorov length scale is
defined as ηk = (ν 3 /)1/4 , where  is the average rate of energy dissipation per unit mass, and ν is the kinematic
viscosity of the fluid. Moin & Mahesh [20] articulated that the smallest resolved length scale in DNS should be of the
order of the Kolmogorov length scale, not lower than or equal; thus, ∆y < 10ηk . Furthermore, according to Batchelor
[10], the smallest resolved thermal length scale in DNS should be of the order of the Obukhoff-Corrsin length scale, ηθ
for P r < 1, or Batchelor scale, λB for P r ≥ 1. For the present study (P r = 0.71), the Obukhoff-Corrsin length scale
is defined as ηθ = ηK P r−3/4 , which is larger than the Kolmogorov length scale. Thus, by resolving the Kolmogorov
length scales, the Obukhoff-Corrsin length scales should also be captured.
+
Figure 2 depicts the wall-normal resolution (∆y + ), the Kolmogorov length scale (ηK ) and the Obukhoff-Corrsin length
scale (ηθ+ ) in local wall units at x/δinlet = 10 and 55. Note that the plots are in a log-log scale. It can be observed that
+
the grid resolution is within the order of the smallest hydrodynamic and thermal length scales (i.e., ∆y + < 10×ηK and
+ +
∆y < 10 × ηθ ), which demonstrates that a sufficient amount of grid points is clustered in the wall-normal direction.
The dashed vertical lines indicate the hydrodynamic boundary layer thickness δ + , non-dimensionalized by the inner
length scale ν/uτ . The Kolmogorov and the Obukhoff-Corrsin length scales exhibit a slight increase in the very strong
FPG zone (x/δinlet = 55), which is consistent with the onset of the quasi-laminarization process (laminarescent stage)
+
and the decrease of the turbulent kinetic energy dissipation. In both streamwise stations, an abrupt increase of ηK
+ +
and ηθ occurs around δ .
The streamwise variation of some boundary layer parameters is shown in figure 3. The momentum and thermal
boundary layer thicknesses, normalized by their inlet values (i.e., δ/δinlet and δT /δT,inlet , respectively), exhibit the
typical linear increasing trend of canonical or ZPG boundary layers in equilibrium flow. As the flow penetrates into
the very strong FPG region, the momentum boundary layer undergoes a shrinking process. Interestingly, there is an
evident change of the δ-slope around x/δ ≈ 34. This steeper decreasing behavior of δ may reveal that the laminarescent
4 Structural Reynolds analogy

Fig. 3: Streamwise variation of boundary layer parameters: velocity/thermal boundary layer thicknesses, skin friction
coefficient and momentum thickness Reynolds number.

stage has initiated. Contrarily, the streamwise variation of δT in the FPG region shows a much slower decreasing rate.
The Reynolds number based on the momentum thickness (i.e., Reθ = U∞ θ/ν) varies almost linearly not only in the
ZPG region (increasing) but also in the FPG region (decreasing). The maximum value of Reθ nearly coincides with
the maximum slope of the skin friction coefficient Cf at the downstream vicinity of the ZPG-FPG intersection, where
Cf = 2(uτ /U∞ )2 . According to Spalart [28], the present flow should have been relaminarized√ as K > 3.0 × 10−6 .
However, note that Cf never achieves the laminar value for sink flows (i.e., Cf = 2.31 K = 0.00462 proposed by
Rosenhead [27]), indicating that the flow is at the earlier stage of relaminarization.
Figure 4 (a) and (b) show snapshots of flow movies based on one iso-surface extracted from the instantaneous stream-
wise velocity (= 0.8U∞,inlet ) and instantaneous temperature (= 0.8T∞ ), respectively. Here, U∞,inlet is the constant
freestream velocity in the ZPG zone (while increasing in the FPG zone) and T∞ is the freestream temperature, which
remains constant outside the thermal boundary layer in the whole domain. These videos are available at [1] and [2],
respectively, and were developed by means of the open-source 3D computer graphics software Blender. The presence
of turbulent bulges is quite evident in the ZPG zone. While in the very strong FPG region, they exhibit a substan-
tial debilitation and penetrate further into the boundary layer, which is attributed to the significant augmentation
of the time-averaged vertical component of the velocity, V , towards the wall. These bulges in the velocity field are
better observed in Figure 4 (c), depicting iso-contours of instantaneous streamwise velocity in the middle spanwise
XY plane of the computational domain. There are valleys of non-turbulent fluid from freestream that penetrate into
the boundary layer and separate the large eddies or bulges, which resemble “islands”. It can be seen how bulges or
large scale motions are tilted toward the wall as the flow moves further into the FPG zone. Additionally, a significant
cessation of entraintment is observed by the end of the computational domain. Conversely, the thermal field seems not
to be similarly affected by the strong flow acceleration. Furthermore, the Reynolds analogy (defined as the similarity
between the momentum and thermal transport) is not fulfilled in the very strong FPG region, caused by the presence
of streamwise pressure gradient, which is a source of dissimilarity between the velocity and temperature field [13].
Whereas a very good affinity can be observed between hydrodynamic and thermal turbulent structures in the ZPG
zone, since streamwise pressure gradient is absent, as seen in figure 5. It was demonstrated in [5] the substantial
increases of the mean vertical velocity V (by more than one order of magnitude with respect to mean vertical velocity
profiles in the ZPG zone) with linear trends in wall units, i.e. V + ∼ −y + . Thus, the mean momentum transport toward
the wall (V ∂U/∂y) or convective transport in the FPG region is clearly observed in the iso-surfaces of instantaneous
streamwise velocity (= 0.8U∞,inlet ) of fig. 5 where inlet velocity values of U in the ZPG region moves closer to the
near wall zone as the flow penetrates deeper into the FPG region.
Figure 6 depicts iso-contours of streamwise velocity fluctuations u0RM S , wall-normal velocity fluctuations vRM
0
S , and
0
thermal fluctuations tRM S in the XY plane and spanwise averaged, which are normalized by the corresponding inlet
freestream parameter, i.e. either velocity or temperature, respectively. While the freestream velocity remains constant
in the ZPG zone, it experiences a significant increase (about 80% by the end of the computational domain) in the
very strong FPG zone given by the prescribed top convergent surface and acceleration parameter K, as discussed in
fig. 3 (a) of Araya and Rodriguez [9]. On the other hand, an isothermal condition was set for the temperature field in
the top surface of the whole domain (at a higher temperature than that of the bottom surface). According to figure 7
(a) maximum local values of u0RM S (when normalized by the inlet freestream velocity U∞,inlet ) remain approximately
constant in the ZPG region (following the typical value of 14% for a turbulent flat plate in low Reynolds numbers),
slightly decrease after prescription of the very strong FPG showing a small dip, to finally exhibit a monotonic linear
Structural Reynolds analogy in laminarescent boundary layers via DNS 5

(a) 0.8U∞,inlet (b) 0.8T∞

(c) XY view

Fig. 4: Iso-surfaces of instantaneous (a) streamwise velocity and (b) temperature. Iso-contours of instantaneous
streamwise velocity in the middle spanwise XY plane (c).

Fig. 5: Top view of iso-surfaces of instantaneous streamwise velocity (= 0.8U∞,inlet , gray) and temperature (=
0.8T∞ , red).

0
increase from x/δinlet ≈ 30 up to 35% by the end of the computational domain. On the contrary, (vRM S )max /U∞,inlet
ratios show reductions of their absolute intensities (up to 20%) due to flow acceleration, see figs. 6 (b) and 7 (a). An
0
analogous decreasing behavior in the order of 23% has been observed in (wRM S )max /U∞,inlet ratios (not shown).
0
The iso-contours of wall-normal velocity fluctuations vRM S in figure 6 (b) depicts an evident thickening of the viscous
sublayer in the FPG region due to the reduction of v 0 in the near wall region.
Sreenivasan [29] states that the turbulent boundary layer usually reacts to acceleration by increased absolute values
of turbulence intensities, particularly from the streamwise component u0 , while this increase might prolong to the
laminarescent region. This is consistent with our observations from figures 6 (a) and 7 (a), where only the streamwise
component of turbulence intensities exhibited increases in the acceleration zone. Clearly, the laminarescent region in
our DNS begins around x/δinlet ≈ 30, and since (u0RM S )max /U∞,inlet does not depict a decreasing trend in fig. 7
(a), this region extends till the end of the computational domain. On the other hand, if (u0RM S )max is normalized by
the local freestream velocity, U∞,local , a subtle linear decreasing behavior is obtained. Badri Narayanan and Ramjee
[24] performed a series of experiments on reverse transition in 2D accelerated incompressible turbulent boundary lay-
ers. In their Experiment No 5, it was prescribed a lengthy zone with constant values of the acceleration parameter
(K = 5.0 × 10−6 ) slightly higher than ours. The (u0RM S )max /U∞,local profile in their Fig. 15 exhibited also a linear
decreasing behavior in the zone of constant K.
It is worth highlighting that velocity fluctuations have been normalized by the inlet freestream velocity; as a con-
sequence, u0RM S has depicted an increasing tendency in the very strong FPG zone. This tendency on u0RM S is the
opposite (so, decreasing) in case of selecting relative or local values of U∞ . Moreover, iso-contours (fig. 6 (c)) and local
maxima (fig. 7 (a)) of t0RM S exhibit almost constant values, even in the very strong FPG zone, which means that
the passive scalar field is nearly unaffected by the presence of strong flow acceleration, whose thermal boundary layer
depicts a very slow shrinking development as seen in fig. 3.
In figure 7 (b), the wall-normal location of fluctuating component maxima in terms of the local boundary layer
thickness (y/δ) is shown. The vertical positions of (u0RM S )max in the ZPG zone are within 8-10% of δlocal . These
relative positions decrease at the beginning of the FPG zone, and finally show a recovery trend from x/δinlet ≈ 30.
These outward displacements of maximum values of the intensity u0 are consistent with experiments by Blackwelder
6 Structural Reynolds analogy

(a) u0RM S /U∞,inlet

0
(b) vRM S /U∞,inlet

(c) t0RM S /T∞,inlet

Fig. 6: Iso-contours of Root Mean Square (RMS) of fluctuating flow components normalized by inlet freestream
values: (a) streamwise velocity fluctuations u0RM S , (b) wall-normal velocity fluctuations vRM
0
S , and (c) thermal
0
fluctuations tRM S .

(a) (b)

Fig. 7: (a) Local maxima of u0RM S , vRM


0 0
S , and tRM S , (b) wall-normal location of maxima in terms of the local
boundary layer thickness.

& Kovasznay [12] in strongly favorable pressure gradient, which reveals the thickening process of the viscous sublayer.
Interestingly, vertical positions of (t0RM S )max depict similar trends as those of (u0RM S )max . The vertical locations of
0
(vRM S )max show some scattering and are roughly at 30% of δlocal in the ZPG zone, reach a minimum of 25% in the
ZPG-FPG intersection (where δlocal is maximum), and finally tend asymptotically to 35% by x/δinlet ≈ 55. The latter
phenomenon could be explained by the substantial shrinking process of the boundary layer thickness in the FPG zone,
while peaks of v 0 persist almost at the same physical vertical location.
Structural Reynolds analogy in laminarescent boundary layers via DNS 7

(a) < u0 v 0 > /U∞,inlet


2

(b) < v 0 t0 > /(U∞ T∞ )inlet

Fig. 8: Iso-contours of (a) Reynolds shear stresses and (b) wall-normal turbulent heat fluxes

Fig. 9: Local maxima of Reynolds shear stresses and wall-normal turbulent heat fluxes.

Figure 8 exhibits iso-contours of Reynolds shear stresses < u0 v 0 > and wall-normal turbulent heat fluxes < v 0 t0 >
in the XY plane and spanwise averaged, where brackets denote time and spanwise averaging. In a similar fashion,
these cross-correlations have been normalized by inlet freestream parameters. Iso-contours of < u0 v 0 > show certain
weakening around 32 < x/δinlet < 44 in the very strong FPG zone. From figure 9, local maxima of < u0 v 0 > are nearly
constant in the ZPG zone, decrease downstream of the FPG setting (for about 14δinlet ’s), and finally recover beyond
2
x/δinlet ≈ 44 towards a value of 0.22% of U∞,inlet by the end of the domain. The latter behavior is more consistent
with the observed increases of u rather than the v 0 trend. In addition, the observed “plateau” in absolute values of
0

< u0 v 0 > in fig. 9 around 32 < x/δinlet < 44 has also been reported by Sreenivasan [29] based on data from [23] (so
called “freezing” of Reynolds shear stresses in outer streamlines). We have reported in [5] a steady decrease of the
Reynolds shear stresses in wall units for the same FPG flow, which is explained by the fact that turbulence stresses
are “diluted” by the increased local mean flow [29]. Figure 9 exhibits a much slower decay of < v 0 t0 >, whose local
< v 0 t0 >max values present a “plateau” in the last 18δinlet ’s. Profiles of the streamwise turbulent heat fluxes < u0 t0 >
or cross-correlation between the streamwise velocity u0 and temperature t0 fluctuations are depicted in figure 10. As
previously done in this study, the local streamwise turbulent heat fluxes are normalized by inlet freestream values:
U∞,inlet and T∞,inlet . Profiles are shown at the following streamwise stations x/δinlet = 10 (ZPG), 30, 40, 50 and 55.
Major outcomes are three-fold: (i) a good collapsing level of all profiles is found in the near wall region, up to 4% of
the boundary layer, (ii) peak values of the cross-correlation < u0 t0 > located around y/δ ≈ 0.09 in the ZPG region
move farther from the wall in the very strong FPG region (≈ 1% of δ), and (iii) these peak values are larger as the
flow moves downstream. The latter is mainly caused by the increasing trend of the component u0 in accelerating flows,
which is consistent with increments of local maxima of u0RM S in the FPG region, as seen in fig. 7.
8 Structural Reynolds analogy

Fig. 10: Profiles of streamwise turbulent heat fluxes < u0 t0 > in outer units.

Figure 11 depicts instantaneous iso-surfaces of quadrant events selected based on a specific threshold value (i.e., Q2 or
ejections and Q4 or sweeps) and cut by a XZ plane at y + =15 with iso-contours of instantaneous streamwise velocity
fluctuations u0 . Regions of ejections clearly dominate over sweeps in the entire domain, with mushroom-like shapes in
the ZPG zone. It is worth to point out that ejections tend to vanish and become highly stretched as the flow accelerates,
particularly in the buffer layer where peaks of u0 and maximum turbulence production occur (see XZ plane at y + =
15). This Q2 reduction in the near wall region due to the very strong FPG condition imposed can be associated with
weakening in bursting rate (and thus in turbulence production). However, ejections are non-negligible by the end of
the computational domain in the inner region (notice Q2 iso-surfaces above low speed streaks or zones of u0 < 0).
This confirms that the flow is situated at a much earlier stage than complete relaminarization (i.e., laminarescense).
On the other hand, Q4 regions show a shallow shape and significantly vanish in the FPG zone.
High/low speed streaks are visualized in Figure 12 (a) by extracting positive/negative values (±0.2U∞,inlet ) of the
instantaneous streamwise velocity fluctuations. Low-speed streaks have the typical wavy shape in the ZPG zone
characterized by upwelling motions of low momentum fluid with streamwise lengths of the order of 4-7δinlet . Whereas,
high-speed streaks penetrate much closer to the near wall region with shorter streamwise lengths (around 3δinlet ).
Turning to the FPG zone, streaks are highly stretched and aligned in the streamwise direction, showing lengths of up to
13δinlet . Figure 12 (b) exhibits positive/negative extracted values (±0.07U∞,inlet ) of the vertical velocity fluctuations,
v 0 . Fluid zones toward the wall (negative v 0 ) depict elongated shapes, which corresponds to Q3 and Q4 or sweep events.
On the contrary, fluid zones away from the wall (positive v 0 ) show a voluminous structure, characterized by Q1 and
Q2 or ejection events. In general, both positive and negative iso-surfaces of v 0 undergo a significant weakening process
as the flow accelerates; however, flow structures with +v 0 seem lesser affected. Instantaneous thermal fluctuations, t0 ,
are depicted in figure 12 (c) by extracting two characteristic iso-surfaces (±0.2T∞ ). A high correlation is observed
between low momentum fluid (negative u0 in fig. 12 (a)) and low temperature fluid (negative t0 ), and viceversa in
the ZPG zone. Whereas this similarity tends to disappear in the very strong FPG zone, associated to the presence
of a source of dissimilarity between the hydrodynamic and passive scalar field: the streamwise pressure gradient.
Furthermore, fluid parcels with low temperature (i.e., t0 < 0) are mostly affected by a streamwise stretching process due
to flow acceleration. Finally, figure 12 (d) shows the instantaneous cross-correlation between the wall-normal velocity
and thermal fluctuations (v 0 t0 or wall-normal turbulent heat fluxes) by extracting two values ±0.007 × (U∞ T∞ ).
As expected, v 0 and t0 are predominantly negatively correlated in the entire domain. The presence of very strong
streamwise FPG provokes a moderate attenuation of the negative wall-normal turbulent heat fluxes, in agreement
with decreases of local maxima of < v 0 t0 > in figure 9.

4 Conclusions

Structural Reynolds analogy is performed in laminarescent boundary layers for the velocity and passive scalar field.
Direct simulation of the continuity, momentum and passive scalar transport equations has been performed in sink-flow
boundary layers. Computation of the Kolmogorov and the Obukhoff-Corrsin length scales has demonstrated the mesh
suitability in capturing the smallest dissipative eddies. The most important conclusions in very strong FPG flows
regarding the velocity/temperature analogy are as follows: (1) streamwise pressure gradient is an important source
of dissimilarity between the momentum and thermal transport, causing a breakdown of the Reynolds analogy, (2)
while evident “signatures” of the very strong FPG have been identified in the velocity field, those “signatures” are
Structural Reynolds analogy in laminarescent boundary layers via DNS 9

Fig. 11: Iso-surfaces of Q2 (green) and Q4 (red) events. Flow in the positive X-direction.

(a) u0 /U∞,inlet (b) v 0 /U∞,inlet

(c) t0 /T∞,inlet (d) v 0 t0 /(U∞ T∞ )inlet

Fig. 12: Iso-surfaces of instantaneous streamwise velocity (u0 ), vertical velocity (v 0 ), thermal (t0 ) fluctuations and
their cross-correlation (v 0 t0 ).

much less obvious in the temperature field; particularly, the thermal boundary layer growth rate (dδT /dx) and thermal
fluctuations do not seem to be strongly affected by the “quasi-laminarization” process, (3) absolute intensities of peaks
of cross-correlation < u0 t0 > increase as the flow moves downstream in the FPG region, while these peaks move farther
from the wall, (4) fluid parcels with low temperature (i.e., t0 < 0) are more altered by the stretching process due to
flow acceleration than those of high temperature streaks (i.e., t0 > 0), and (5) the survival of ejections after almost
40δinlet streamwise length from the very strong FPG prescription suggests that flow reversion is a very lengthy and
slow process, particularly, for a passive scalar field.
10 Structural Reynolds analogy

Acknowledgements GT acknowledges the Puerto Rico Louis Stokes Alliance for Minority Program. GA acknowledges AFOSR grant
FA9550-17-1-0051 and NSF-CBET grant #1512393. Computational resources were supplied by XSEDE (project #TG-CTS170006).

References

1. https://youtu.be/w8NB_-4pvG0.
2. https://youtu.be/KBpH8ixD0Ng.
3. R. J. Adrian and Z.C Liu. Observation of vortex packets in Direct Numerical Simulation of fully turbulent channel flow. Journal
of Visualization, 5:9–19, 2002.
4. T. Antonia, H. Abe, and H. Kawamura. Analogy between velocity and scalar fields in a turbulent channel flow. Journal of Fluid
Mechanics, 628:241–268, 2009.
5. G. Araya, C. Castillo, and F. Hussain. The log behaviour of the Reynolds shear stress in accelerating turbulent boundary layers.
Journal of Fluid Mechanics, 775:189–200, 2015.
6. G. Araya and L. Castillo. DNS of turbulent thermal boundary layers up to Reθ = 2300. Int. Journal of Heat and Mass Transfer,
64:4003–4019, 2012.
7. G. Araya, L. Castillo, C. Meneveau, and K. Jansen. A dynamic multi-scale approach for turbulent inflow boundary conditions in
spatially evolving flows. Journal of Fluid Mechanics, 670:518–605, 2011.
8. G. Araya, S. Leonardi, and L. Castillo. Steady and time-periodic blowing/suction perturbations in a turbulent channel flow. Physica
D: Nonlinear Phenomena, 240(1):59–77, 2011.
9. G. Araya and D. Rodriguez. Visualization and assessment of turbulent coherent structures in laminarescent boundary layers. J. of
Visualization, https://doi.org/10.1007/s12650-017-0460-4, 2017.
10. G. K. Batchelor. Small-scale variation of convected quantities like temperature in turbulent fluid. part 1. general discussion and
the case of small conductivity. Journal of Fluid Mechanics, 5:113–133, 1959.
11. D. M. Bell and J. H. Ferziger. Turbulent boundary layer DNS with passive scalars. In In Near-Wall Turbulent Flows, edited by R.
M. C. So, C. G. Speziale, and B. E. Launder, pages 327–336. Addison-Wesley, 1993.
12. B. F. Blackwell, W. M. Kays, and Moffat R. J. The turbulent boundary layer on a porous plate: An experimental study of the heat
transfer behavior with adverse pressure gradients. Technical Report HMT-16, Stanford University, August 1972.
13. G. Araya and L. Castillo. DNS of turbulent thermal boundary layers subjected to adverse pressure gradients. Physics of Fluids,
page 095107, 2013.
14. M. Guala, S. E. Hommema, and R. J. Adrian. Large-scale and very-large-scale motions in turbulent pipe flow. Journal of Fluid
Mechanics, 554:521–542, 2006.
15. S.J. Kline, W.C. Reynolds, F.A. Schraub, and P.W. Runstadler. The structure of turbulent boundary layers. Journal of Fluid
Mechanics, 30:741–773, 1967.
16. H. Kong, H. Choi, and J.S. Lee. Direct numerical simulation of turbulent thermal boundary layers. Physics of Fluids, 12(10):2555–
2568, 2000.
17. Q. Li, P. Schlatter, L. Brandt, and D. Henningson. DNS of a spatially developing turbulent boundary layer with passive scalar
transport. Int. J. Heat Fluid Flow, 30:916–929, 2009.
18. T.S. Lund, X. Wu, and K.D. Squires. Generation of turbulent inflow data for spatially-developing boundary layer simulations.
Journal of Computational Physics, 140(2):233–258, 1998.
19. Y. Mito and N. Kasagi. Quasi-coherent turbulent structures in a channel with an oscillatory deformed wall. Journal of Visualization,
1:129–129, 1998.
20. P. Moin and K. Mahesh. Direct numerical simulation: a tool in turbulence research. Annu. Rev. Fluid Mech., 30:539–578, 1998.
21. R. Narasimha. Relaminarization-magnetohydrodynamic and otherwise. AIAA Progress in Astronautics and Aeronautics, 84:30–53,
1983.
22. R. Narasimha and K.R. Sreenivasan. Relaminarization in highly accelerated turbulent boundary layers. Journal of Fluid Mechanics,
61:417–447, 1973.
23. M. A. Badri Narayanan, S. Rajagopalan, and R. Narashima. Some experimental investigations on the fine structure of turbulence.
Rep 74FM 15, Dept. Aero. Eng., Ind. Inst. Sci., Bangalore., 1974.
24. M. A. Badri Narayanan and V. Ramjee. On the criteria for reverse transition in a two-dimensional boundary layer flow. Journal
of Fluid Mechanics, 35:225–241, 1969.
25. U. Piomelli and J. Yuan. Numerical simulations of spatially developing, accelerating boundary layers. Phys. Fluids, 25:101304,
2013.
26. S. Pirozzoli, M. Bernardini, and P. Orlandi. Passive scalars in turbulent channel flow at high Reynolds number. Journal of Fluid
Mechanics, 788:614–639, 2016.
27. L. Rosenhead. Laminar boundary layers. Oxford University Press, London., 1963.
28. P. Spalart. Numerical study of sink-flow boundary layers. Journal of Fluid Mechanics, 172:307–328, 1986.
29. K.R. Sreenivasan. Laminarescent, relaminarizing and retransitional flows. Acta Mech., 44:1–48, 1982.
30. I. Tiselj. Tracking of large-scale structures in turbulent channel with direct numerical simulation of low Prandtl number passive
scalar. Physics of Fluids, 26:125111, 2014.
31. Z. Warhaft. Passive scalars in turbulent flows. Ann. Rev. Fluid Mechanics, 32:203–240, 2000.
32. P. Wu, X. & Moin. Transitional and turbulent boundary layer with heat transfer. Physics of Fluids, 22:085105, 2010.
33. X. Wu. Inflow turbulence generation methods. Ann. Rev. Fluid Mechanics, 49:23–49, 2017.
34. H. Zhao, A. Wei, K. Luo, and J. Fan. Direct Numerical Simulation of turbulent boundary layer with heat transfer. International
Journal of Heat and Mass Transfer, 99:10–19, 2016.

You might also like