scispace - formally typeset
Open AccessProceedings ArticleDOI

An electrothermally-aware full-chip substrate temperature gradient evaluation methodology for leakage dominant technologies with implications for power estimation and hot-spot management

Reads0
Chats0
TLDR
In this paper, an accurate chip-level electrothermally-aware methodology for spatial silicon substrate temperature estimation is presented, which self-consistently incorporates various electrothermal couplings arising mainly due to the strong dependence of sub-threshold leakage on temperature and also employs an accurate package thermal model, to account for inhomogeneous layers and non-cubic structure.
Abstract
As CMOS technology scales into the nanometer regime, power dissipation and associated thermal concerns in high-performance ICs due to on-chip hot-spots and thermal gradients are beginning to impact VLSI design. Moreover, elevated substrate (junction or die) temperature strongly influences IC performance, reliability, and packaging/cooling cost. Hence, accurate estimation of substrate thermal profiles is critical. This paper presents an accurate chip-level electrothermally-aware methodology for spatial silicon substrate temperature estimation. The methodology self-consistently incorporates various electrothermal couplings arising mainly due to the strong dependence of subthreshold leakage on temperature and also employs an accurate package thermal model, to account for inhomogeneous layers and non-cubic structure, which are not considered in traditional methods. The proposed methodology becomes increasingly effective as technology scales due to increasing leakage. Furthermore, it is shown that considering realistic package thermal models not only improves the accuracy of estimating temperature distribution but also has significant implications for power estimation and hot-spot management.

read more

Content maybe subject to copyright    Report

An Electrothermally-Aware Full-Chip Substrate Temperature Gradient
Evaluation Methodology for Leakage Dominant Technologies with
Implications for Power Estimation and Hot-Spot Management
Sheng-Chih Lin and Kaustav Banerjee
Department of Electrical and Computer Engineering, University of California, Santa Barbara, CA 93106
{sclin, kaustav}@ece.ucsb.edu
ABSTRACT
As CMOS technology scales into the nanometer regime, power
dissipation and associated thermal concerns in high-performance ICs
due to on-chip hot-spots and thermal gradients are beginning to
impact VLSI design. Moreover, elevated substrate (junction or die)
temperature strongly influences IC performance, reliability, and
packaging/cooling cost. Hence, accurate estimation of substrate
thermal profiles is critical. This paper presents an accurate chip-level
electrothermally-aware methodology for spatial silicon substrate
temperature estimation. The methodology self-consistently
incorporates various electrothermal couplings arising mainly due to
the strong dependence of subthreshold leakage on temperature and
also employs an accurate package thermal model, to account for
inhomogeneous layers and non-cubic structure, which are not
considered in traditional methods. The proposed methodology
becomes increasingly effective as technology scales due to increasing
leakage. Furthermore, it is shown that considering realistic package
thermal models not only improves the accuracy of estimating
temperature distribution but also has significant implications for
power estimation and hot-spot management.
1. INTRODUCTION
While continued scaling of CMOS technologies provides
substantial benefits in transistor density and circuit performance,
increasing chip power consumption and power densities are rapidly
leading to thermal management concerns. Moreover, highly integrated
circuits including System-on-chips (SoCs) with different functional
blocks, blocks with different activity rates (for example, logic vs.
memory) and clock/power gating techniques essentially create non-
uniform temperature distributions across the chip substrate [1]. The
regions with higher temperature are commonly referred to as hot-spots.
Hot-spots simultaneously lead to temperature gradients that affect
performance [2] (including delay and timing) and reliability among a
host of other issues, and also result in a general over-design in high-
performance microprocessor packaging and cooling solutions. These
thermal problems have now been identified and projected as major
challenges for future IC design by leading semiconductor
manufacturers and by the International Technology Roadmap for
Semiconductors (ITRS) [3].
Permission to make digital or hard copies of all or part of this work for
personal or classroom use is granted without fee provided that copies are
not made or distributed for profit or commercial advantage and that
copies bear this notice and the full citation on the first page. To copy
otherwise, to republish, to post on servers or to redistribute to lists,
requires prior specific permission and/or a fee.
ICCAD'06, November 5-9, 2006, San Jose, CA
Copyright 2006 ACM 1-59593-389-1/06/0011...$5.00
1.1 Implications of Substrate Temperature Rise and
Non-Uniform Thermal Profile
Elevated substrate temperature is widely known to have a strong
impact on the lifetime of devices and back-end reliability, including
interconnects under “field,” “accelerated testing” and “burn-in”
conditions. Due to non-uniform power distribution, temperature in the
local hot-spot regions can easily exceed the maximum reliability limit
and increase the risk of damaging the device and interconnect (since
major reliability mechanisms, including electromigration (EM), time-
dependent dielectric breakdown (TDDB), and negative bias
temperature instability (NBTI) have a strong dependence on
temperature) even with advanced thermal management technologies
[4]. Moreover, due to the increase in the number of interconnect levels
and introducing low-k dielectric materials with poor thermal
conductivity, thermal problems have become even worse [5-7].
High temperature not only leads to the onset and acceleration of
reliability problems at the device and interconnect level but also
impacts on circuit- and system-level metrics. Increased temperature
deteriorates circuit performance by degrading device carrier mobility
and increasing interconnect metal resistivity. This, in turn, will impact
physical design issues including Power/Ground integrity [8] and
placement and routing schemes [9-11]. At the system-level, thermal
management (packaging and cooling) solutions are also affected by
substrate temperature because they have to meet the maximum heat-
flux requirements at the silicon-package interface [12].
1.2 Measurement and Modeling of Substrate Thermal
Profile: Prior Work
Although thermal infrared imaging system is used in the
industry for acquiring chip thermal profiles, it is not a useful tool
during the design process, but is merely a verification technique after
chip fabrication and packaging. Also, for typical high-performance
microprocessors, the thermal profile obtained by the infrared imaging
system does not accurately reflect the realistic substrate thermal
profile of an operating microprocessor with a sophisticated packaging
structure. Similarly, integrated thermal sensors are commonly
employed to ensure that hot-spots do not exceed the specified
maximum temperature criteria in high-performance microprocessors.
However, only a rudimentary thermal profile with low resolution can
be detected by these integrated thermal sensors, since the number of
sensors that can be integrated into a chip is limited by routing and
pin-out constraints.
In order to predict the thermal gradient as well as the
temperature profile of high-performance ICs, especially
microprocessors, several methodologies have been developed to
perform a full-chip thermal analysis. In [13, 14], a chip-level
temperature profile is generated by a numerical finite difference
approach incorporating temperature dependent device models and
lumped R-C network models. This approach solves the partial
differential equations (PDE) of heat transfer by direct matrix
factorization, which becomes complicated for large scale problems.
Different thermal simulation algorithms have been proposed for
568

improving computation efficiency. A chip-level 2-D and 3-D thermal
simulator is presented in [15, 16]. Instead of direct matrix solving, the
simulator solves the similar heat diffusion PDE by performing the
Alternating Direction Implicit method with higher efficiency. In [17],
multigrid (MG) method, along with coarsening grid process, is
presented to reduce the memory usage for computation. In [18], a
combination of Green’s function method and transformation is
proposed for highly efficient steady-state thermal analysis. A full-
chip thermal simulation methodology using pre-calculated constant
power dissipation at the functional block level is proposed in [19]. In
[20], analysis for a full-chip and cooling system thermal model is
presented. However, all these analysis are mainly focused on either
improving algorithms for solving heat transfer equations or
accelerating the computational efficiency for temperature estimations
(improving the simulation runtime within a range of minute). Also,
these analyses are all based on a cubic (unrealistic) thermal model for
the entire chip and the packaging stack-up, which in turn,
compromises the accuracy of thermal estimation because the
unrealistic package thermal model neglects the effect of heat
spreading at different packaging layers.
Most importantly, due to technology scaling and parameter
variations [21, 22], including non-uniform dopant distribution in the
channel region of the transistors [23], leakage power dissipation,
which is dominated by subthreshold leakage for high-performance
ICs, becomes a significant component of total chip power
consumption. The subthreshold leakage is exponentially dependent
on temperature [24] and exacerbates with technology scaling [25].
Also, the increase in total chip power consumption causes higher
substrate temperature, which further increases the subthreshold
leakage, thereby creating a strong feedback loop leading to various
electrothermal couplings between power, temperature, operating
frequency and supply voltage [26]. All prior substrate temperature
profile simulation methods not only employ an overly simplistic
package thermal model [13-20], but also ignore these electrothermal
couplings that are an inseparable aspect of nanometer scale chip
operation. Hence, unlike previous works that target only
computational efficiency, we propose a full-chip thermal analysis
methodology that incorporates these electrothermal couplings, as well
as a realistic package thermal model to improve the accuracy of the
substrate thermal profile estimation and the methodology is
implemented via one of the widely-used efficient algorithms for
solving the heat transfer diffusion equations.
The rest of the paper is organized as follows. Details of various
electrothermal couplings between power dissipation, operating
frequency and substrate temperature are described in Section 2. In
Section 3, we formulate a realistic package thermal model and
incorporate electrothermal couplings into heat partial differential
equations. In Section 4, the implementation and discussion of
proposed methodology are outlined. Impact of the realistic package
thermal model formulation and implications of the electrothermally-
aware methodology for power estimation and thermal management
are discussed in Section 5. Finally, concluding remarks are made in
Section 6.
2. ELECTROTHERMAL COUPLINGS
Chip power dissipation at the nanometer scale has two major
components: switching power and leakage power dissipation (The
short-circuit component is relatively small and therefore we neglect it
throughout this paper). The switching power consumption increases
with the chip frequency and supply voltage. Moreover, the
performance itself is dependent on temperature. Increase in substrate
temperature will decrease the transistor drive current due to the
reduction in carrier mobility (although the threshold voltage decreases
at higher operating temperature and partially offsets the performance
degradation resulting from the lower carrier mobility, the transistor
drive current still decreases at higher operating temperatures) as
shown in Fig. 1(a).
As mentioned earlier, subthreshold leakage, the main leakage
contributor, is highly temperature sensitive (Fig. 1(b)). Moreover, due
to technology scaling and parameter variations, leakage power
dissipation becomes a major contributor to total chip power
consumption [25]. The increase in total chip power consumption
causes higher on-chip temperature, which further increases the
subthreshold leakage. Therefore, a strong feedback loop leading to
various electrothermal couplings occurs [26]. Fig. 2 illustrates various
electrothermal couplings between performance (frequency), power
dissipation, supply voltage, threshold voltage, and substrate
temperature. Hence, in order to accurately estimate power dissipation
and resulting substrate temperature profile, various electrothermal
couplings must be embedded into full-chip thermal modeling and
analysis.
Figure 1. (a) Drive (drain) current (b) Off-state leakage current for NMOSFET
(45 nm and 90nm technology nodes) as a function of operating temperature.
Figure 2. Schematic view of electrothermal couplings between different design
parameters.
3. FULL-CHIP REALISTIC PACKAGE THERMAL
MODEL
We consider a realistic microprocessor package structure, Flip-
Chip Land Grid Array (FC-LGA) and a socket that interfaces with the
printed-circuit board (PCB) as shown in Fig. 3.
The microprocessor die is mounted on a package substrate. An
integrated heat spreader (IHS) is attached to the package substrate and
microprocessor die. The IHS spreads the non-uniform heat from the
die to the top of the IHS, and it improves the heat flux from a smaller
die area to a larger surface and serves as the mating surface for a
heatsink. Since the surface of these three major components (die, IHS
and heatsink) are never smooth enough to have perfect contact, they
are bonded together with a thermal interface material (TIM) applied
between them. The thermal interface material improves the poor
thermal conductivity caused by surface roughness (conductivity of
569

TIM is much larger than air) and thus, enhances the overall thermal
performance of the stack-up packaging and cooling mechanisms.
Moreover, a minor heat transfer path exists from the die to the
printed-circuit board (PCB) mainly corresponding to the
interconnect/dielectric layers and I/O pads. The thermal conductivity
of this path (from substrate to the printed-circuit board) is normally
several orders of magnitude smaller than that of the major heat
transfer path [27]. Therefore, we neglect this path throughout this
paper because of the small fraction of heat it can transfer.
Figure 3. Sketch of microprocessor package assembly. (drawing not to scale)
Heat is considered to be a form of energy that can be transferred
as a result of temperature difference by three different modes:
conduction, convection, and radiation. However, since radiative heat
losses are negligible for a packaged chip, we only consider conduction
and convection in this paper.
In a three-dimensional system, heat conduction and convection
can be quantified by Fourier’s law and Newton’s law of cooling as
shown in (1) and (2) respectively.
T
kA
n
()
surface
hA T T
where the negative sign indicates that the heat transfer will be a
positive quantity in the direction of decreasing temperature (i.e.,
temperature gradient, T, is negative), based on the second law of
thermodynamics. The surface area normal to the direction of heat
conduction is represented by A. The outward direction normal to the
surface A is represented by n. The quantity, T, is the temperature
distribution of the material. The thermal conductivity of the material
is denoted by k and is a measure of the ability of the material to
conduct heat. The thermal conductivity varies with temperature and
depends on material characteristics [28]. Here, we use the thermal
conductivity at the average temperature and treat it as a constant in all
calculations. Also, we consider the thermal conductivity of each
packaging layer to be isotropic. h denotes the convection heat transfer
coefficient. T
surface
is the surface temperature, and T
is the
environmental temperature sufficiently far from the surface.
Equivalent thermal resistances can be derived from (3) and (4)
below under different scenarios (conduction and convection). By the
duality of electrical and thermal quantities, power dissipation (heat
flow), temperature, and thermal resistance are analogous to current
flow, voltage, and electrical resistance, respectively. The equivalent
thermal model can be established by a thermal resistance network to
represent these inhomogeneous packaging material layers, and solving
the voltages of the network gives the temperature distribution of all
the layers. However, for large scale problems, this approach becomes
complicated when both computational efficiency and profile
resolution are of importance.
θ
=
conduction
L
kA
1
convection
hA
θ
=
In order to improve the thermal performance of the major heat
transfer path, typically, a larger dimension of heat spreader and
heatsink is used. In practice, the area of heat spreader and heatsink is
at least 9X and 30X larger than the area of a die, respectively.
Traditional chip-level thermal analyses employ a cubic thermal model
for simplicity as shown in Fig. 4(a). Although the resolution of the die
region and the computation efficiency could be improved, this
unrealistic package thermal model underestimates the lateral heat
spreading due to large packaging layers. Fig. 4(b) illustrates the
relative dimensions of realistic packaging layers, which we considered
in the proposed methodology. Note that not only are they different
materials with different thermal properties but also their dimensions
with respect to the silicon die will significantly influence the heat
transfer as well as the substrate thermal profile.
Figure 4. Side view of thermal packaging stack-up. The layout, power density
distribution, and dimension of the die are identical for both packaging cases.
The thickness of different layers and the dimension of the layers are not drawn
to scale. (a) Cubic packaging model (b) Realistic packaging model indicating
different dimensions for each layer.
The silicon die is the main source of heat generation. Heat can be
exchanged and transferred by conduction within the entire packaging
stack-up and convection at the surface of the heatsink. The
fundamental physics of heat transfer in a chip substrate is governed by
the following three-dimensional heat conduction equation and subject
to heat convection as the boundary condition [28]:
[]
(, ,,) (, ,,) (, ,,) (, ,,)
ρ
=∇ +
p
C Txyzt kxyzt Txyxt gxyzt
t
[]
(,,,) (,,,) (,,,)
=−
amb
i
kxyzt Txyzt hTxyzt T
n
where ρ is the density of the material (kg/m
3
), C
p
is the specific heat
of material (J/kg°C), T is the temperature (°C), k is the thermal
conductivity of the material (W/m°C), g is the internal heat generation
(W/m
3
), n
i
is the outward direction normal to the boundary surface, h
is the heat transfer coefficient (W/m
2
°C), and T
amb
is the temperature
of the ambient air surrounding the package measured at a specified
distance sufficiently far away from the surface of the entire package.
As mentioned earlier, we consider each discretized layer to be
isotropic and homogeneous. Therefore, we use a constant thermal
conductivity within one layer, and the temperature of the entire
structure will be modeled by rewriting the partial differential equation
570

and boundary condition as (7) and (8) where temperature (T) is a
function of the position (x, y ,z) and time (t).
222
222
()
ρρ

∂∂∂
=+++

∂∂

pp
Tk TTTp
t Cxyz C
[]
=−
amb
i
Th
TT
nk
Figure 5. Sketch of the discretization of the thermal packaging stack-up. Each
node (circle) in the figure represents a discretized cell with a temperature (T).
Each discretized cell has six adjacent cells connected by edges (lines).
Relationships between two adjacent cells are governed by (7) or (8) depending
on heat transfer mechanisms. Effective thermal conductivity of cells between
two adjacent layers (darker nodes) can be determined by (9) since the
dimensions of each discretized cell are equal (i.e., dx = dy = dz).
As discussed in the previous section, various electrothermal
couplings need to be considered and incorporated into the thermal
model and analysis. Therefore, the parameter p in equation (7) is a
function of temperature, time and the position within the die. It
represents a modified parameter of the constant quantity g in (5) for
the internal heat generation. Since there are electrothermal couplings
between power dissipation, operating frequency and die temperature,
the quantity p is not a constant value like g and will be evaluated in a
self-consistent manner at each iteration step.
We discretize the entire thermal packaging stack-up
(inhomogeneous packaging material layers) based on a typical
microprocessor package structure according to its physical
dimensions. Relationships between discretized cells are governed by
the heat partial differential equations and boundary conditions shown
in (7) and (8). Physical thermal parameters, such as thermal
conductivity, density, and specific heat of different layers, depend on
material properties. Note that the dimensions of a discretized cell are
chosen to be equal (i.e., dx = dy = dz). Thus effective thermal
conductivity (k
eff
) of cells between two adjacent layers, as represented
by a darker node in Fig. 5 between layer 1 and layer 2, can be simply
determined by (9).
12
211
()=+
eff
kkk
where k
1
, k
2
represent the thermal conductivity of material 1 and 2,
respectively. k
eff
is the effective thermal conductivity between these
two layers. Since thermal interface material (TIM) is applied between
two different layers (Fig. 3) to reduce the contact resistance caused by
surface roughness, we assume there is a perfect thermal contact
between the TIM layer and the other material.
Due to the presence of complex geometry and the complicated
boundary condition, the silicon junction temperature profile can not be
solved analytically. However, a numerical solution can be found by
finite difference approaches and approximation schemes. In the next
section, the self-consistent electrothermally-aware methodology for
estimating substrate temperature profile is proposed and implemented.
4. IMPLEMENTATION OF THE
ELECTROTHERMALLY-AWARE METHODOLOGY
Several numerical approaches exist in the literature for solving
partial differential equations [28, 29]. Here we use the Alternating-
Direction-Implicit method [30, 31] because it is a widely used method
for the efficient numerical solution of parabolic partial differential
equations in multiple spatial variables. The advantage of applying this
method is that we are able to transfer a multiple dimensional parabolic
partial differential equation into a succession of one-dimensional
problems. Therefore, no large scale matrix has to be computed and it
is easy to be implemented. Thus, we use the alternating direction
implicit method as the core algorithm to solve the heat partial
differential equations for achieving higher computation efficiency. It
is important to note that although other computationally efficient
methods exist, but choosing any one of them over the others does not
affect the core results of our proposed methodology. The key aspect of
our self-consistent approach is that it inherently generates a more
accurate power profile (Fig. 6), which can then be used to generate a
temperature profile using any computationally efficient PDE solvers.
Figure 6. Key aspect of the electrothermally-aware methodology. Due to the
strong interdependence of temperature and leakage power, temperature at the
center block (T
0
) is not simply a function of the power dissipation within and
adjacent to the center block as per traditional analysis. Nominal power
distribution map should be updated self-consistently with the temperature
evaluation.
The overview of the proposed electrothermally-aware substrate
temperature gradient evaluation methodology is illustrated in Fig. 7.
The chip (target simulation domain) is partitioned into a mesh
according to the information provided by the layout geometry
(position) and power distribution map. Nominal power distribution
(including switching and leakage power dissipation) for each
functional block according to its activity factor is used as initial values
depending on the circuit implementation and technology nodes.
Physical parameters such as specific heat, thermal conductivity and
heat transfer coefficient depend on given packaging material
properties and applied cooling techniques. The full-chip realistic
packaging model is incorporated and comprehends both vertical and
lateral heat transfer paths. Boundary conditions are determined by the
operating environment. The simulator uses the layout geometry,
power distribution, boundary conditions, and physical thermal
parameters as initial values to formulate parabolic partial differential
energy equations and then solves these equations in a self-consistent
manner using the Alternating-Direction-Implicit method for every
mesh element. The algorithm converts a multiple dimensional
parabolic partial differential equation into a succession of one-
dimensional linear problems. The electrothermal couplings are also
embedded in the core of the simulator that simultaneously estimates
temperature dependent quantities for each simulation step. Once the
difference of the temperature evaluation between two steps is within a
571

certain range (e.g. 0.01°C), the evaluation stops and the steady-state
temperature profile is obtained. However, if the temperature exceeds
the maximum criteria (defined by reliability constraints) for certain
extreme cases due to poor packaging/cooling solutions or high power
dissipation, the evaluation stops and thermal runaway will be reported.
Figure 7. Overview of the electrothermally-aware silicon temperature profile
simulator.
We further implemented the proposed simulator on a PC (3.06
GHz Pentium 4 processor, 1 GB memory) using C++ language. A
microprocessor design with die size of 10 mm × 10 mm (discretized
into 100×100 grids) and with power densities per functional block is
shown in Fig. 8. Note that we consider the power dissipation of each
block calculated according to each block’s nominal switching power,
leakage power, and activity factor. The nominal total power
consumption of the chip at nominal temperature (25°C) is 96 W
(nominal active power = 93.1 W, leakage power = 2.9 W). The short-
circuit component is relatively small; therefore we neglect it for
simplicity. The physical and thermal properties of the packaging
layers are similar to a practical package of a high-performance
microprocessor.
The proposed methodology has significant implications for
various temperature-dependent effects because of the accurate
substrate temperature profile it can generate. For instance,
temperature increase in the local areas (hot-spots) tends to accelerate
device and interconnect failure mechanisms that are strongly
temperature dependent. Also, signal integrity analysis under non-
uniform substrate temperature [32] requires an accurate substrate
temperature estimation. At the circuit level, buffer insertion and gate
sizing (physical design process) can be made thermally aware
because interconnect and gate delays are strongly dependent on
temperature [33]. Besides synthesis, placement, and routing
algorithms, power-grid analysis can be made thermally aware to
ensure acceptable voltage-drop levels in the presence of significant
chip substrate temperature gradients [8, 10].
In the next section, a comparison between different scenarios is
shown to highlight the importance and the impact of employing
electrothermal couplings and realistic package thermal models for
substrate thermal gradient estimation. The proposed methodology
inherently provides an accurate estimation of power dissipation in
leakage dominant CMOS technologies. Moreover, implications for
IC-cooling and thermal (hot-spot) management for nanometer scale
ICs are also presented and discussed.
0 2000 4000 6000 8000 10000
0
2000
4000
6000
8000
10000
0.6
0.8
1
1.2
1.4
1.6
1.8
Figure 8. Functional block layout of a test chip. Power densities associated
with functional blocks are also shown. The circle indicates a region where
functional blocks have highest power density. The triangle indicates the
functional blocks that have higher leakage power dissipation than all other
blocks.
5. SIMULATION RESULTS AND IMPLICATIONS
5.1 IMPACT OF ELECTROTHERMAL COUPLINGS
First we demonstrate the importance of incorporating
electrothermal couplings for estimating substrate temperature profile.
Although the results are specific to the above mentioned IC, the
conclusions are more generic. It can be observed that there is a region
indicated by a circle in Fig. 8 where blocks within this region have
highest power density. In addition, there is a region indicated by a
triangle in Fig. 8 where blocks have 10X leakage power dissipation
with respect to the values of all other functional blocks.
The substrate temperature profile shown in Fig. 9 is generated
using traditional thermal simulator without considering electrothermal
couplings. The highest temperature (hot-spot) is around 64.2°C and
located in a region with highest power density (indicated by a circle in
Fig. 8). However, based on the same packaging and system cooling
condition (model shown in Fig. 4(b)), a different substrate
temperature profile (Fig. 10) is obtained by the proposed
electrothermally-aware simulator. From the temperature profile, two
hot-spots can be observed--one in the region with highest power
density and the other in the region with higher percentage of leakage
power. Unlike traditional evaluation, the highest temperature is
around 63.8°C and is located in a region with higher percentage of
leakage power (indicated by a triangle in Fig. 8).
As explained in Section 2, a region with higher switching power
density does not necessarily yield a higher temperature due to the
various electrothermal couplings. It is easy to observe the impact of
electrothermal couplings on substrate temperature evaluation by
comparing Fig. 9 and Fig. 10. The substrate temperature profile
obtained by electrothermally-aware evaluation shows an additional
hot-spot and also a different temperature distribution. The traditional
estimation is clearly misleading in terms of hot-spot count, location,
and the overall spatial temperature profile as it neglects the
electrothermal couplings between power dissipation and temperature.
572

Figures
Citations
More filters
Journal ArticleDOI

Phd by thesis

TL;DR: In this paper, a sedimentological core and petrographic characterisation of samples from eleven boreholes from the Lower Carboniferous of Bowland Basin (Northwest England) is presented.
Journal ArticleDOI

Reliability- and process-variation aware design of integrated circuits.

TL;DR: A broad review the literature for Reliability- and Process-variation aware VLSI design shows a re-emergence of the topic as a core area of active research and is likely to be a part of any reliability qualification protocol for future technology generations.
Proceedings ArticleDOI

Thermal via allocation for 3D ICs considering temporally and spatially variant thermal power

TL;DR: This paper develops an accurate and efficient thermal-via allocation considering the temporally and spatially variant thermal-power, and is solved by a sequential quadratic programming algorithm with use of sensitivities from the macromodel.
Journal ArticleDOI

Power Blurring: Fast Static and Transient Thermal Analysis Method for Packaged Integrated Circuits and Power Devices

TL;DR: A new method, called power blurring (PB), is developed for calculating temperature distributions using a matrix convolution technique in analogy with image blurring, which provides much more accurate temperature profiles with reduced computation time.
References
More filters
Journal ArticleDOI

Phd by thesis

TL;DR: In this paper, a sedimentological core and petrographic characterisation of samples from eleven boreholes from the Lower Carboniferous of Bowland Basin (Northwest England) is presented.
Book

Fundamentals of Modern VLSI Devices

Yuan Taur, +1 more
TL;DR: In this article, the authors highlight the intricate interdependencies and subtle tradeoffs between various practically important device parameters, and also provide an in-depth discussion of device scaling and scaling limits of CMOS and bipolar devices.
Proceedings ArticleDOI

Parameter variations and impact on circuits and microarchitecture

TL;DR: Process, voltage and temperature variations; and their impact on circuit and microarchitecture; and possible solutions to reduce the impact of parameter variations and to achieve higher frequency bins are presented.
Related Papers (5)