Download PDF
ads:
Riemann solutions of balance systems with
phase change for thermal flow in porous media.
Adviser: Dan Marchesin
Co-adviser: Johannes Bruining
Student: Wanderson Jos
´
e Lambert
May, 2006
1
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
Contents
1 Introduction 9
2 General Formulation 15
2.1 Physical situations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.1 Accumulations and fluxes for a physical situation . . . . . . . . . 21
2.1.2 Riemann problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.1.3 Wave Sequences and Riemann Solution . . . . . . . . . . . . . . . 24
2.2 Characteristic speeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3 Shock waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3.1 Shocks and the Rankine-Hugoniot condition . . . . . . . . . . . . 29
2.3.2 The Rankine-Hugoniot locus and shock waves . . . . . . . . . . . 30
2.3.3 Extension of the Bethe-Wendroff theorem . . . . . . . . . . . . . . 37
2.4 Regular RH locus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.4.1 Small shocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.4.2 Other degeneracies of RH locus . . . . . . . . . . . . . . . . . . . . 47
2.4.3 The sign of u
+
on regular Rankine-Hugoniot loci . . . . . . . . . . 47
3 Mathematical modelling of thermal oil recovery with distillation 50
3.1 Physical and Mathematical Models . . . . . . . . . . . . . . . . . . . . . . 51
3.1.1 Balance Equations. . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.1.2 Reduced system of equations . . . . . . . . . . . . . . . . . . . . . 55
3.2 Physical situations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.2.1 Three phases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.2.2 Two phases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.2.3 One phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4 The Riemann solution for the injection of steam and nitrogen 63
4.1 Physical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.2 The model equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2
ads:
4.3 Physical situations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.3.1 Single-phase gaseous situation - spg ................. 67
4.3.2 Two-phase situation - tp ........................ 68
4.3.3 Single-phase liquid situation - spl ................... 69
4.3.4 Primary, secondary and trivial variables . . . . . . . . . . . . . . . 69
4.4 General theory of Riemann Solutions . . . . . . . . . . . . . . . . . . . . . 70
4.4.1 Characteristic speeds in each physical situation . . . . . . . . . . 70
4.4.2 Shock waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.4.3 Bifurcation loci . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.4.4 Admissibility of shocks . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.5 Elementary waves in the single-phase gaseous situation . . . . . . . . . 81
4.5.1 Characteristic speed analysis . . . . . . . . . . . . . . . . . . . . . 81
4.5.2 Shocks and contact discontinuities . . . . . . . . . . . . . . . . . . 83
4.6 Contact discontinuities in the single-phase liquid situation . . . . . . . . 84
4.7 Elementary waves in the two-phase situation . . . . . . . . . . . . . . . . 85
4.7.1 Characteristic speed analysis . . . . . . . . . . . . . . . . . . . . . 85
4.7.2 Shocks in the two-phase situation . . . . . . . . . . . . . . . . . . 90
4.8 Shocks between regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.8.1 Shock between the gaseous and the two-phase situations . . . . . 92
4.8.2 Shock between the two-phase and the liquid situations . . . . . . 94
4.9 The Riemann solution for geothermal energy recovery . . . . . . . . . . . 96
4.9.1 Subdivision of tp ............................. 96
4.9.2 The Riemann solution . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.9.3 V
L
in III and IV .............................100
5 Riemann solution for steam and water flow 101
5.1 Mathematical and Physical model . . . . . . . . . . . . . . . . . . . . . . 103
5.1.1 The model equations . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.1.2 Physical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.2 Regions under thermodynamical equilibrium . . . . . . . . . . . . . . . . 105
5.3 Equations in conservative form . . . . . . . . . . . . . . . . . . . . . . . . 106
5.4 Elementary waves under thermodynamical equilibrium . . . . . . . . . . 106
5.4.1 Steam region - sr .............................107
5.4.2 Boiling region - br ............................109
5.4.3 Water region - wr ............................110
5.5 Shocks between regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.5.1 Water Evaporation Shock . . . . . . . . . . . . . . . . . . . . . . . 111
5.5.2 Vaporization Shock . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.5.3 Condensation Shock . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.5.4 Steam condensation front . . . . . . . . . . . . . . . . . . . . . . . 116
5.6 The Riemann Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.6.1 Riemann Problem A . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.6.2 Riemann Problem B . . . . . . . . . . . . . . . . . . . . . . . . . . 122
5.6.3 Riemann Problem C . . . . . . . . . . . . . . . . . . . . . . . . . . 126
3
5.6.4 Riemann Problem D . . . . . . . . . . . . . . . . . . . . . . . . . . 137
5.6.5 Riemann Problem E . . . . . . . . . . . . . . . . . . . . . . . . . . 139
6 Riemann solution for problem of Chapter 4 for V
L
in III and IV. 145
6.1 V
L
in L
6
......................................145
6.2 V
L
in L
5
......................................146
7 Summary and Conclusions 148
A Physical quantities; symbols and values for the Nitrogen Problem 154
A.1 Temperature dependent properties of steam and water . . . . . . . . . . 156
A.2 Constitutive relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
B Physical quantities; symbols and values for the steam injection 158
B.1 Temperature dependent properties of steam and water . . . . . . . . . . 158
B.2 Constitutive relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
B.3 Approximation for T
+ M
ρ
g
c
g
(T)/C
r
in the sr ................159
B.3.1 Rarefaction wave behavior . . . . . . . . . . . . . . . . . . . . . . . 161
C Applications 163
C.1 Hyperbolic waves in a physical situation . . . . . . . . . . . . . . . . . . . 163
C.2 Numerical Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
C.2.1 General Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 167
C.2.2 Numerical Method for the System . . . . . . . . . . . . . . . . . . 167
C.3 Travelling waves and viscous profiles. . . . . . . . . . . . . . . . . . . . . 168
4
`
As duas mulheres da minha vida na ordem em que
apareceram: Ivete e Milene.
Abstract
This work encompasses two types of results. First we present a new general the-
ory which deals with Riemann solution for a large class of balance equations. This
class of equations has interest because it can be applied to model thermal flow with
mass interchange between phases in porous media, with important applications in oil
recovery.
As applications of this theory, we present three models of steam injection in a
horizontal porous media. The systems of equations are based on mass balance, energy
conservation andDarcy law of force. We neglect compressibility, heat conductivity and
capillarity effects.
In first example we consider steam/water/oil flow in porous medium. We only
prepare the formalism for this class of equations.
In second example we consider steam/water/nitrogen flow into a porous medium.
We develop the general theory for a 4
× 4 system of balance equations. We solve the
Riemann problem with application to recovery of geothermal energy. A rarefaction
evaporation wave is observed.
In third example we consider the steam/water injection into a porous medium. We
completely solve the Riemann problem associated with this model. We obtain a rich
class of bifurcations in the solutions. A new type of shock, the evaporation shock, is
identified in the Riemann solution.
Keywords: Porous medium, steamdrive, Riemann solution, balance equa-
tions, multiphase flow, distillation
Resumo
Este trabalho abrange dois tipos de resultados. Primeiramente n
´
os apresentamos
uma nova teoria geral que trata da soluc¸
˜
ao de Riemann para uma grande classe de
equac¸
˜
oes de balanc¸o. Esta classe de equac¸
˜
oes tem interesse porque pode ser aplicada
ao modelamento de fluxos t
´
ermicos em meios porosos com transfer
ˆ
encia de massa
entre fases, com importantes aplicac¸
˜
oes na recuperac¸
˜
ao de
´
oleo.
Como aplicac¸
˜
oes desta teoria, n
´
os apresentamos tr
ˆ
s modelos de injec¸
˜
ao de vapor
em meios porosos horizontais. Os sistemas de equac¸
˜
oes s
˜
ao baseados no balanc¸o de
massa, na conservac¸
˜
ao de energia e na lei de Darcy da forc¸a. N
´
os negligenciamos a
compressibilidade, a condutividade do calor e efeitos de capilaridade.
No primeiro modelo, n
´
os consideramos fluxo de vapor/
´
agua/
´
oleo no meio poroso.
N
´
os somente preparamos o formalismo para esta classe de equac¸
˜
oes para resolvermos
o problema de Riemann como um trabalho futuro.
No segundo modelo, n
´
os consideramos fluxo de vapor/
´
agua/nitrog
ˆ
enio num meio
poroso. N
´
os desenvolvemos uma teoria geral para um sistema 4
× 4 de equac¸
˜
oes de
balanc¸o. Resolvemos o problema de Riemann com aplicac¸
˜
ao
`
a recuperac¸
˜
ao de energia
geot
´
ermica. Uma onda de rarefac¸
˜
ao de evaporac¸
˜
ao
´
e observada.
No terceiro exemplo, consideramos a injec¸
˜
ao de vapor/
´
agua em um meio poroso.
N
´
os resolvemos completamente o problema de Riemann associado com este modelo.
Obtemos uma rica classe de bifurcac¸
˜
oes para as soluc¸
˜
oes. Um novo tipo de choque, o
choque da evaporac¸
˜
ao, identificado na soluc¸
˜
ao de Riemann.
Palavras-chaves: meios porosos, injec¸
˜
ao de vapor, soluc¸
˜
ao de Riemann,
equac¸
˜
oes de balanc¸o, fluxo multif
´
asico, destilac¸
˜
ao.
Agradecimentos
Agradecimentos prescindem de ordem. Aos que eu agradec¸o, n
˜
ao h
´
a aqueles que
tenham sido mais importantes. Todos foram imprescind
´
ıveis nos momentos certos,
pois uma tese n
˜
ao se erige em quatro anos. Uma tese comp
˜
oe-se de gr
˜
aos: dos estudos
anteriores, de cada momento dispensado, de cada impress
˜
ao da vida. Colaboram uns
mais, outros menos: desde os professores da inf
ˆ
ancia at
´
e o orientador.
Portanto s
˜
ao muitos a serem agradecidos e se eu, injustamente, esquecer de algu
´
em,
desculpem-me.
Agradec¸o sobretudo a Deus que em sua eterna sabedoria soube esconder a matem
´
atica
nos improv
´
aveis rec
ˆ
onditos do Universo.
`
A matem
´
atica pela poesia e arte que se amalgamam entre seus furtivos segredos.
Ao meu orientador pela quase infinita paci
ˆ
encia com a qual me orientou na tese,
no ingl
ˆ
es e na vida.
Ao Professor Johannes Bruining que me deu boas noc¸
˜
oes de termodin
ˆ
amica e
procurou sempre sanar minhas impertinentes d
´
uvidas.
Aos meus antigos orientadores da UNICAMP: Petr
ˆ
onio e Marcelo. O primeiro me
enveredou para a matem
´
atica e o segundo para as leis de conservac¸
˜
ao.
`
A banca, por ter me ajudado a melhorar a vers
˜
ao final.
Ao IMPA, aos seus funcion
´
arios e ao incompar
´
avel suporte de pesquisa que a mim
foi propiciado.
Aos professores que me fizeram enxergar o mundo com outros olhos, a todos eles.
Aos funcion
´
arios do laborat
´
orio de fluidos pelo apoio, pelo suporte, pela ajuda ou
simplesmente pela amizade.
Aos novos amigos que adquiri no Rio (de alguma forma ou de outra todos atrav
´
es
do IMPA) e que ajudaram a amainar a inexor
´
avel saudade: da namorada, da fam
´
ılia
e dos velhos amigos.
`
A minha fam
´
ılia pelo apoio. A esse micro-cosmo que
´
e meu suporte. Em especial
`
a minha m
˜
ae,
`
a tia Francinete, aos meus irm
˜
aos e irm
˜
as, cunhadas e cunhados e as
crianc¸as que muitas vezes atrapalharam meus estudos e sempre me faziam participar
de suas travessuras.
`
A Milene, uma impressionante mulher que me preencheu de amor. Que mesmo
longe, afugentava meus medos e me afagava com sua terna m
˜
ao em minhas lembranc¸as.
Agradec¸o a ela pela paci
ˆ
encia e pelas brigas...pois sempre era doce o retorno.
Agradec¸o aos meus amigos de Cambu
´
ı. Que s
˜
ao apenas amigos mas poderiam ter
sido irm
˜
aos. Agradec¸o por cada riso, pelas discuss
˜
oes, pelo compartilhar de id
´
eias e
da partilha das inquietaes terrenas que por vezes experimentamos.
Por fim agradec¸o ao CNPq que parcialmente financiou esta pesquisa e a ANP pelo
suporte financeiro e a oportunidade de desenvolver este trabalho.
CHAPTER 1
Introduction
A particular class of balance equations are the conservation laws. The physical
interpretation of these laws is that for any limited spatial domain
D R
d
the change
the total amount of any quantity (for example, mass, momentum, energy) is due ex-
clusively to the flow of this quantity through the boundary of the domain.
In the Mathematics literature, a system of conservation laws in one spatial dimen-
sion is usually defined as:
U
t
+ F
x
= 0, (1.1)
where U
R
m
represents a state vector field of dependent variables. The func-
tion F :
= F(U) :=(F
1
(U), F
2
(U),···, F
n
(U))
T
: −→ R
m
is the flux of such quantities,
generically, it is a
C
2
function. The derivatives U
t
= U/t and F
x
(U)=F/x are
interpreted in the distribution sense, so that
(1.1) is regarded to be in the weak form;
this is important because of the presence of shocks.
An important problem associated to Eq.
(1.1) appeared in the work of Riemann
in the XIX century. He was interested in the evolution of two gases with different
pressures separated by a thin membrane within a tube. Suddenly, the membrane
is broken. He analyzed the evolution of the gas flow. In this model he observed
rarefaction waves and discontinuous solutions, which are called shocks nowadays.
This work originated the study of an important class of Cauchy problems, the so
called Riemann problem, governed by equations of type
(1.1) with initial data of the
form:
U
(x, t = 0) :=
U
0
, if x < 0
U
1
, if x > 0
. (1.2)
Riemann problems have the important property that solutions are invariant under
the scaling x
ax, t at, for a > 0. For non-linear problems (1.1), they play
the role of fundamental solutions, in a way analogous to sines and cosines for linear
problems.
9
Conservation laws have been used to model a variety of physical phenomena (such
as fluid dynamics, flow in porous medium, road traffic, electromagnetism, etc...) and
therefore the theory for this class of equations is well developed, see for example
[60,20,59,45] and for numerical methods [46,36,34].
In most applications it is useful to utilize a more general form of conservation
laws:
G
t
+ F
x
= 0, (1.3)
where G :
= G(U) :=(G
1
(U), G
2
(U),···, G
n
(U))
T
: −→ R
m
also are C
2
. G is the
accumulation term; the equations are interpreted in the distribution sense.
In classical models of flow in porous medium, the system of equations can be writ-
ten in the form, see
[12]:
G
t
+(uF)
x
= 0, (1.4)
where u is the total or Darcy speed. However, incompressibility is often assumed in
these models, which implies that u is constant spatially. As a result, in such flows
(1.4) reduces to the form (1.3).
In many problems, such as transport of hot fluids and gases undergoing mass gain,
loss or transfer, flows involving chemical reactions such as combustion balance laws
generalizing
(1.3) are required to describe the flow. The classical form of balance laws
is:
G
t
+ F
x
= Q. (1.5)
Here Q :
= Q(U) :=(Q
1
(U), Q
2
(U),···, Q
m
(U))
T
: −→ R
m
are C
2
functions that
represent sources, sinks or transference of the physical quantities. Although there
are many applications for this class of equations, its theory is not well consolidated.
In our work, we are interested in models for flow in porous media, for such flows
Q represents mass transfer of chemical components between phases with no net gain
or loss of component mass, as well as conservation of total energy. The variable u
representing the Darcy speed also appears, but only in particular way within the flux
term:
G
t
+
(
uF
)
x
= Q. (1.6)
The variables in
(1.6) are V∈ R
m
and u ∈R. In this model, the variable u is not
constant, generically.
System of type
(1.6) model thermal compositional flows in porous media. The
variable u is called speed because this is its interpretation in many applications; the
pair
(V, u) in R
n+1
is called state variable. G and F are the vector-valued functions
G =(G
1
,G
2
,···,G
m+1
)
T
: −→ R
m+1
and F =(F
1
,F
2
,···,F
m+1
)
T
: −→ R
m+1
,
where u
F
i
is the flux for the conserved quantity G
i
and G
i
/t is the correspond-
ing accumulation term, for i
= 1,2,···,m + 1. On the right hand side in Q =
(
Q
1
, Q
2
,···, Q
m+1
)
T
: −→ R
m+1
, the first m terms represent mass transfer and
Q
m+1
represents energy conservation. Generically for a thermally isolated system,
the first m equations represent mass balance for different chemical components in
different phases and the
(m + 1)-th equation represents the conservation of total en-
ergy, which usually can be represented by setting Q
m+1
= 0. The functions G, F
and Q are continuous in the whole domain , but later we will see that they are
10
only piecewise C
2
. The equations are supplemented by relationships expressing local
thermodynamic equilibrium in each of the
C
2
parts.
The solution
V(x, t) and u(x, t) for x R and t R
+
needs to be determined. The
terms Q often generate fast variations in the solution due thermodynamic processes;
this is another reason to regard the solution as distributions and
(1.6) must be taken
in the weak sense. Despite the fact that the variable u does not appear in the accumu-
lation term, but only in the flux term and despite the presence of the transfer term Q
formally breaking scale invariance, we are able to solve the complete Riemann prob-
lem associated to Eq.
(1.6). In fact, a general theory to deal this class of equations is
proposed in this work, extending recent works of Bruining et. al.
[5, 6, 8, 9, 10, 11].
This theory encompasses and generalizes the bifurcation phenomena for systems that
change type discovered over the last two decades, see
[24,25,26].
In Chapter 2, we define the physical situations, in which one or several phases or
chemical components are missing, determining the expression of the prevailing local
thermodynamic equilibrium. In each of these physical situations, by combining the
original equations the balance system
(1.6) reduces to simpler systems of conserva-
tion laws with fewer equations of the form:
t
G
(V)+
x
uF
(V)=0, (1.7)
where the transfer terms Q has cancelled out. In a physical situation governed by
(1.7), the corresponding set of variables V is a subset of the set of variables V, while
F and G are obtained from
F and G; they have n + 1 components, n m. The physical
domain in each physical situation is denoted by
j
for j = I, II, III, ··· and it is a sub-
set of . There are three groups of variables: the basic variables V, called “primary
variables”; the variable u, called “secondary variable” because it is obtained from the
primary variables; and the “trivial variables” are constant or can be recovered from
other variables in a simple way. Notice that the numbers of primary, secondary and
trivial variables always add up to m
+ 1.
In
[15], Colombo et. al studied a problem with phase transitions in 2 × 2 systems
of type
(1.1). They considered that the physical domain is formed by two disjoint
sub-domains
X
1
and X
2
. The domains are physical situations, which Colombo called
phases. The phase transition is the jump in the solution with left and right states
belonging to different phases. Our model is physically more adequate because it in-
cludes also infinitesimally small phase transitions, so that
X
1
adjoins X
2
, as they
often are do in actual physics.
The class of equations studied in this work was motivated by practical problems of
steam injection in porous medium. Mathematical models for steam injection can be
found in the works of Coats
[13], Marx et. al. [50], Lake et. al. [23] and Pope et. al.
[54]. The rigorous analysis we are pursuing started in the works of Bruining et. al.
[5, 6, 8, 9, 10, 11]. Following these initial ideas, we focus our interest in Riemann
problems for variants of steam injection. Techniques based on steam injection have
much interest because they are used in oil industry. For half a century, direct of cyclic
steam injection have been applied as a method for enhancing recovery of oil, mainly
11
for heavier oil. Heavy oil at low temperatures is highly viscous and so recovery is
very difficult. Steam injection is used to increase the oil temperature, improving
recovery both because the oil viscosity is lowered and because the volatility of light
oil is increased, concentrating it in certain regions and allowing it to act as a liquid
solvent that displaces the heavy oil, see
[4 ] and [11].
Steam flood is part of the so called thermal methods, see
[55]. Thermal methods
are divided into two large groups. One group encompasses the techniques in which
heat is injected into the porous medium, a category encompassing steam injection.
In other group heat is generated locally in the porous medium, usually by means of
combustion. In oil recovery, combustion utilizes a small part of the oil in the reser-
voir. Air is injected and the oxygen participates in the chemical reactions. As steam
injection, combustion phenomena are modelled by balance equations, see
[1 ], and the
works of Coats
[14], Crookston et. al. [16]. Modern theory of conservation laws was
used for the study of combustion by Mota et. al.
[27, 28, 29, 30, 31, 32], focusing on
the study of the internal structures of combustion front which is a classical problem
in engineering, see Souza et. al.
[62].
In Chapter 3, we present the mathematical and physical model for flow of water,
steam, volatile and dead oil. Dead oil is a heavy oil with negligible vapor pressure
while volatile oil is a light oil with non-negligible vapor pressure. A partial Riemann
solution of this problem was found by Bruining and Marchesin
[11].
In the last decades, steam injection has been adapted for clean-up of organic con-
taminants from the subsurface, the so called groundwater remediation. The organic
pollutants are referred to as non-aqueous phase liquids; they are divided into two
large groups: DNAPL and LNAPL. The DNAPLs are denser than water; examples of
DNAPL are chlorinated solvents. It is very difficult to decontaminate soils containing
these substances because they can enter deeply underground. The LNAPLs are the
substances less dense than water, such as gasoline and oil.
Traditional remediation of contaminated sites consists of groundwater extraction
and clean up of the drained liquid; this technique is called pump-and-treat, see
[21],
[49] and references therein. The material is treated and discharged in another place.
This method is very expensive and time-consuming, for example, in the Visalia test
site, in California, polluted with creosote (a NAPL fluid), the estimated costs for
pumping and treating are US$26,000 per gallon of creosote and the expected time
for this procedure is 3250 years. It is inefficient to clean up sites contaminated with
DNAPLs (dense non aqueous phase liquids). So removal of contaminants with steam
is considered as an alternative; for the same site in Visalia, the estimated cost per
gallon of creosote removed with steam is US$130 and the time to clean up is approxi-
mately 3 years, see
[17].
In
[18,19], Davis described the mechanism of steam injection. For clean up, steam
is injected in the soil by using wells. Initially the steam heats the formation around
the wells. The steam condenses as the latent heat of vaporization of water is trans-
ferred from the steam to the region around the wells. As more steam is injected, the
hot water moves into the formation, pushing the water initially present in the rock,
which is at the ambient temperature. When the porous medium at the injection re-
12
gion has absorbed enough heat to reach the temperature of the injected steam, steam
itself actually enters the medium, pushing the cold water and the bank of condensed
steam in front of it. When these hot liquids reach the region that contains the volatile
contaminant, the contaminant is displaced. Since the temperature of the volatile con-
taminant increases part of this contaminant evaporates originating a gaseous flow
with organic components.
Notice that in this model there appear several physical situations and that there
is mass transfer between the phases. The total energy is conserved, a fact that is
modelled by using the conservation law for energy. The conservation of mass in the
flow together with Darcy’s law for porous media have motivated the introduction of
equations of type
(1.6). In chapter 2, we propose a general formalism for solving this
class of equations. We consider the different physical situations as simple connected
subsets state of
R
m
. Physical situations are adjacent in state space. There exists a
linear map E that takes the system of balance laws
(1.6) into a system of conservation
laws
(1.7) associated to each physical situation. After this elimination, further sim-
plifications using thermodynamical constraints may occur in each physical situation
is described by a set of variables V that is a subset of the set of variables
V. Of course,
each of the systems of type
(1.7) has fewer equations than the complete system (1.6).
We are interested in the Riemann-Goursat problem where x
= 0 is the injection
point and the data in
(1.2) for x < 0 is replaced by the injection boundary data at
x
= 0. It is well known that the Eqs. (1.1), (1.3) with initial data of (1.2) exhibits
solutions that are constant along lines x
/t, i.e., self-similar solutions. By physical
considerations, we assume that in our model rarefaction waves occur only within each
physical situation, where Eq.
(1.6) reduces to equations of type (1.7). Since there is
fast transfer of mass in the thin space between the physical situations, we propose
shocks linking them. In Sections
[2.2] and 2.3, we prove that the solution states are
constant along the lines x
/t. We also show that we can obtain the rarefaction and
shock waves first in the space of primary variables, and that the secondary variable
u can be recovered in terms of the primary variables.
In Proposition 2.3.1 we summarize the theory showing that the Riemann problem
can be completely obtained in the space of primary variables. Moreover, if a sequence
of waves and states solves the Riemann problem in the primary variables, for a given
u
L
> 0 (or u
R
> 0), then it is also a solution for any other u
L
> 0 (or u
R
> 0), i. e., if
the Darcy speed u
L
in the initial data is modified while V
L
and V
R
are kept fixed, the
Darcy speeds along the Riemann solutions are rescaled in the
(x, t) plane, while the
wave sequences and the Riemann solution in
V(x, t) are kept unaltered. As tools, we
prove three generalization of the Triple Shock Rule
[25]. These rules are important to
compare the speed of shocks of different families and obtain the Riemann solution.
In Section 2.3.3 we generalize the Bethe-Wendroff theorem for shocks between
physical situations for equations of type
(1.6). As in the classical case, extrema of
shock speeds are related to shock/rarefaction resonance. This theorem is very im-
portant for the identification of bifurcation loci in the solution. These structures are
important because the Riemann problem changes when the left (or right) data is pre-
scribed in different regions in the physical domain relatively to these structures.
13
In Section 2.4, we define the notion of regular Rankine-Hugoniot (RH) locus, which
are curves that parametrize shocks. For the class of equations
(1.6), we define the
notion of strictly hyperbolic physical situation and show a generalization of the Lax
theorem for shocks within a physical situation, Proposition 2.4.1. In Section 2.4.3,we
prove that the speed u does not change sign in connected branches of the RH locus.
In Appendix C, we show some applications of the general theory. In Appendix
C.1, we prove that the formalism for the class of equations
(1.6) can be applied to
the simpler classical equation
(1.3). In Appendix C.2, we obtain an efficient quasi-
implicity numerical scheme for equations
(1.7) without knowing the variable u.In
Appendix C.3, we show that the viscous profile for equations
(1.6) can be studied only
in the space of primary variables V, without knowing the speed u.
In Chapter 3, we present a physical model for steam and gaseous volatile oil injec-
tion into a horizontal, linear porous medium filled with oil and water. We neglect com-
pressibility, heat conductivity and capillarity effects and present a physical model for
oil/water/steam injection is based on the mass balance and energy conservation equa-
tions as well as on Darcy’s. We present the main physical definitions and summarize
the equations of the model. This system of equations is a representative example of
flow in porous media with mass interchange between phases (in the absence of grav-
itational effects) Eq.
(1.6). The Riemann solution was partially solved in [11]. Our
purpose is to show how this model fits within the formalism.
In Chapter 4, using the model proposed by Bruining et. al.
[6 ], we formulate a
system that describes the flow of nitrogen, steam and water in a porous medium.
The physical model for nitrogen/steam/water injection is based on the mass balance
and energy conservation equations. We describe the formalism for solving Riemann
problem associated to the a 4
×4 system of balance equations. As example of Riemann
solution, we show an application for the recovery of geothermal energy. Injecting
water and nitrogen into a hot porous rock, the water in the mixture evaporates and
rock thermal heat can be recovered even if the rock is not very hot. A new fact is the
presence of a rarefaction wave associated to evaporation in the two-phase situation.
In Chapter 5, we obtain the solutions for the basic one-dimensional profiles that
appear in the clean up problem or in recovery of geothermal energy. We consider the
injection of a mixture of steam and water in several proportions into a porous rock
filled with a different mixture of water and steam. We describe completely all possible
solutions of the Riemann problem. We find several types of shock between regions
and a rich structure of bifurcations. A new type of shock, the evaporation shock, is
identified in the Riemann solution. This model generalizes the work of Bruining et.
al.
[6 ], where the condensation shock appeared.
Each chapter can be read separately in this thesis. We put in each chapter a
brief introduction explaining the content of each chapter, so that they can be read
separately in this thesis. More precisely, Chapter 3 requires Chapter 2. Chapters 4
and 5 can be read on their own.
Much remains to be studied for models of type
(1.6) as discussed in the conclusion,
Chapter 7.
14
CHAPTER 2
General Formulation
Mass interchange between different phases occurs typically in multiphase flows in
porous media. Such flows are not represented by conservation laws or hyperbolic
equations both because they have source terms and because they are not evolutionary
in all the variables. In the absence of gravitational effects, such flows can be modelled
by systems of m
+ 1 equations:
t
G(V)+
x
u
F(V)=Q(V) , (2.1)
where
V∈ R
m
and u ∈R. The variable u is called speed because this is its
interpretation in many applications; the pair
(V, u) in R
n+1
is called state variable.
G and F represent the vector-valued functions G =(G
1
,G
2
,···,G
m+1
)
T
: −→ R
m+1
and F =(F
1
,F
2
,···,F
m+1
)
T
: −→ R
m+1
, where uF
i
is the flux for the conserved
quantity
G
i
and G
i
/t is the corresponding accumulation term, for i = 1,2,···,m + 1.
On the right hand side Q
=(Q
1
, Q
2
,···, Q
m+1
)
T
: −→ R
m+1
, the first m components
represent mass transfer and Q
m+1
represents energy transfer due to phase change.
Generically for a thermally isolated system, the m first equations represent mass bal-
ance for different chemical components in different phases and the
(m + 1)-th equa-
tion represents the conservation of total energy, which usually can be represented by
setting Q
m+1
= 0. The functions G, F and Q are continuous in the whole domain ,
but later we will see that they are only piecewise smooth.
The solution
V(x, t) and u(x, t) for x R and t R
+
needs to be determined.
The terms Q often generate fast variations in the solution, so that in such cases the
solution may be regarded as distributions and the equation
(2.1) must be taken in the
weak sense. This equation has an important feature: the variable u does not appear
in the accumulation term, but only in the flux term.
Generically, the equations
(2.1) model flows where there are phase changes giving
rise to mass exchange between different phases. Such flows encompass particular
15
physical situations, where one or several phases or chemical components are missing.
The balance system
(2.1) reduces in each particular physical situation to simpler
systems of conservation laws with fewer equations of the form:
t
G
(V)+
x
uF
(V)=0, (2.2)
where the source term Q is absent. In each physical situation, the corresponding set
of variables V is a subset of the set of variables
V; F and G are obtained from F and
G; they have n + 1 components, n m.
Typically, the laws of thermodynamics play a central role in models
(2.1). Each
physical situation where the system of balance equations
(2.1) reduces to a system of
type
(2.2) is actually a physical situation under local thermodynamical equilibrium
or quasi-equilibrium and the system
(2.2) is a system under local thermodynamical
equilibrium or quasi-equilibrium: this equilibrium is enforced by thermodynamic re-
lationships between the quantities in V. There are three groups of variables in each
physical situation. Generically, when physical changes occur in waves under ther-
modynamical equilibrium, the solution of the system
(2.1) and (2.2) are the same;
however, when there is only local thermodynamical equilibrium, the solution of
(2.1)
is an approximation of the ( 2.2). In a future work, we intend to understand how de-
viations in thermodynamical equilibrium interfere in the approximation of
(2.1). The
basic variables V, called “primary variables”; the variable u, called “secondary vari-
able” because it is obtained from the primary variables; and the “trivial variables” are
constant or they can be recovered from other variables in a simple way. Notice that
the number of primary, secondary and trivial variables always add up to m
+ 1.
2.1 Physical situations
We define the physical situation
j
as a subset of that describes the j-th physical
situation. The Roman number index I, II, III,
··· indexes the physical situation; we
represent the set of indices by
PS = {I, II, III, ···}. We denote the points of
j
by
V
j
, see Examples 2.1.1 and 2.1.2 below.
The sets
j
satisfy:
j
and =
j
j
. (2.3)
The sets
P
j
of primary state variables are obtained from
j
, one subset for each
physical situation. The primary variables correspond to the unknowns V
j
of a system
of conservation laws of the form
(2.2). The number of primary variables in P
j
is
indicated by n
(j). The primary variables in each P
j
are denoted by V
j
following the
notation in
(2.2) to indicate that they depend on the physical situation j. For each
physical situation, there are accumulation and flux functions G and F satisfying an
equation of the form
(2.2); G and F are obtained from G and F. To indicate that
16
(G, F) depends also on j, we utilize the index j, so the functions are written as G
j
=
(
G
j
1
, G
j
2
,···, G
j
n
(j)+1
)
T
and F
j
=(F
j
1
, F
j
2
,···, F
j
n
(j)+1
)
T
.
We denote the trivial variables by
V
j
and the space of trivial variables by P
j
.
Notice that in each physical situation
j
, the trivial variables are uniquely de-
termined by the primary variables or solely by the physical situation, so there is a
one-to-one correspondence between
P
j
and
j
, which is denoted by i
j
; its inverse i
j
is defined uniquely, which is the restriction of
j
to P
j
.
Consider a point
V
j
with primary coordinates V
j
∈P
j
, the restriction i
j
satisfies:
i
j
:
j
−→ P
j
, i
j
(V
j
)=V
j
,. (2.4)
Notice that this restriction is a projection of
V
j
into P
j
.
The reverse is a map from
P
j
to V
j
, which is:
i
j
: P
j
−→
j
, i
j
(V
j
)=V
j
. (2.5)
We define the restriction of
G and F to
j
by R
j
G
:
j
−→ R
m+1
and R
j
F
:
j
−→
R
m+1
, respectively. For V
j
j
, they satisfy:
R
j
G
(V
j
)=G(V
j
) and R
j
F
(V
j
)=F(V
j
). (2.6)
The original
G and F are continuous in , but they are not differentiable. However
we assume that the restrictions are smooth, specifically, they belong to
C
2
(
j
,R
m+1
).
We will see below that the functions
(G
j
, F
j
) for the j-th conservation system can
be obtained as a linear combination of
R
j
G
and R
j
F
, so the pairs (G
j
, F
j
) belong to
C
2
(
j
,R
m+1
).
Because we are solving problems in one spatial dimension, certain pairs of physical
situations may occur contiguously. Let
j
and
k
with P
j
and P
k
represent the
primary variables in one of these pairs, with
j
on the left of
k
in (x, t) space; we
use
() for j and (+) for k. Whenever there is a non-empty intersection between a
pair
j
and
k
, we denote it by
jk
,i.e.,
jk
=
j
k
. The elements of
jk
are
denoted by
V
jk
.
In the space of primary variables, this intersection is denoted by Π
jk
and is given
by:
Π
jk
= i
j
(
jk
)=i
k
(
jk
).
Definition 2.1.1. Let
V∈
k
k
. We say that V is a private point of some
j
for
j
∈PS, if there exists no other l ∈PS, such that V∈
l
.
Definition 2.1.2. Let
V∈
k
k
. We say that V is a frontier point of some
j
for
j
∈PS, if there exists at least another l ∈PS, such that V∈
l
.
17
Remark 2.1.1. Notice that if V is a private point, then we can perform the projection
in
P
j
for some j ∈PS, i.e., there exists a V ∈P
j
, such that V = i
j
(V); we call such
V too private point. If
V is a frontier point, then in the applications it can belong to
different physical situations depending on the necessity, but there always exists at least
a j
∈PSand a V ∈P
j
such that V = i
j
(V
j
); we call such V too frontier point.
From the continuity of the accumulation and flux functions
F and G in we
obtain:
Lemma 2.1.1. For all
V
jk
jk
, R
j
G
(V
jk
)=R
k
G
(V
jk
) and R
j
F
(V
jk
)=R
k
F
(V
jk
).
It is important to define the notion of open sets in the topology induced by
j
and
P
j
relatively to R
n( j)
. Generically we are interested in solving the system of equations
to obtain the primary variables in each physical situation, so it is useful to define the
induced topology of
j
from the induced topology of P
j
and the inverse function i
j
,
given by
(2.5). For each P R
n( j)
and
> 0, the open balls are the sets of x R
n( j)
that satisfy || x P||
n( j)
<
, denoted by B
(P). The norm || · || is the euclidian norm
of
R
m
, with m > n( j) and || · ||
n( j)
is the induced norm in R
n( j)
.
Definition 2.1.3. The euclidian norm in the induced topology
j
or P
j
is the
restriction of
|| · || in each space, which are denoted by ||·||
j
and ||·||
P
j
.
Definition 2.1.4. Let V
∈P
j
, the open balls in the induced topology P
j
are B
j
=
P
j
B
(V), and the open balls in the induced topology
j
are i
j
(B
j
) where i
j
and
i
j
are defined in (2.4) and (2.5).
Definition 2.1.5. Let D
⊂P
j
. We say that D is an open subset in the space of
primary variables, if all V
D there exists an
> 0 such that B
j
(V) D. The subset
i
j
(D) is an open subset of
j
.
Definition 2.1.6. Let V
∈P
j
. We say that V is an internal point of P
j
, if there exists
an
> 0 such that B
j
(V) ⊂P
j
. The other points are called boundary points of P
j
.If
V is an internal (boundary) point for a
P
j
, then i
j
(V) is an internal (boundary) point
for
j
.
Remark 2.1.2. Notice that the frontier points are boundary points, however, there are
boundary points that are private points. The points that are boundary and frontier
points are called exterior boundary.
Since the primary variable spaces are projections of the
j
for j ∈PS, it is not
possible to consider the union of different primary variable spaces
P
j
. However, we
can consider the union of some
j
. Since we are interested in physical phenomena
we consider a connected union of the variable spaces
k
, which is denoted by
k
k
.
Remark 2.1.3. Since each physical situation contains its frontiers, it is closed in .
18
Example 2.1.1.
In Chapter 5, we study the problem of steam injection into a porous medium,
proposed originally by Bruining et. al. and partially solved in
[6 ]. In this model
diffusive, capillarity and heat conductivity effects are ignored. The mass balance
equations for liquid water, steam and energy conservation in terms of enthalpies are
written as:
t
ϕρ
W
s
w
+
x
u
ρ
W
f
w
= q
g−→ a,w
, (2.7)
t
ϕρ
g
s
g
+
x
u
ρ
g
f
g
= q
g−→ a,w
, (2.8)
t
ρ
r
h
r
+
ϕ
(
ρ
W
h
w
s
w
+
ρ
g
h
g
s
g
)
+
x
u
(
ρ
W
h
w
f
w
+
ρ
g
h
g
f
g
)=0. (2.9)
Here
ϕ
is the rock porosity assumed to be constant; s
w
and s
g
are, respectively, the
water saturation and steam saturation; u is the Darcy velocity; f
w
and f
g
are the
fractional flow functions for water and steam;
ρ
r
and
ρ
W
are the constant rock and
water densities. The steam density
ρ
g
is a function of the temperature T; h
r
, h
w
and
h
g
are the rock, water and steam enthalpies per unit mass; these properties depend on
temperature (see Appendix B); and q
g−→ a,w
is the water mass transfer term or steam
condensation term. All variables are defined in Appendix B.
19 There are three physical situations: steam situation, labelled by I where there
is only steam above the fixed water boiling temperature; the boiling situation, labelled
by II, where water and steam coexist at boiling temperature; the water situation,
labelled by III, where there is only water below boiling temperature, see Figure 2.1.
In the steam situation the space of primary and trivial variables are
P
I
= {T} and
P
I
= {s
w
= 0} with
I
= {s
w
= 0,T}; the Darcy speed u is an unknown (secondary
variable). In the boiling situation,
P
II
= {s
w
}, P
II
= {T
b
} and
II
= {S
w
, T
b
},
where T
b
is the boiling temperature at the prevailing temperature; the Darcy speed
u is constant (secondary variable) and the flow is governed by the Buckley-Leverett
equation. In the water situation,
P
III
= {T}, P
III
= {s
w
= 1} and
III
= {s
w
= 1,T};
the Darcy speed u is constant (secondary variable).
Example 2.1.2.
In Chapter 4, we study flow of nitrogen and steam in a porous medium with water,
proposed originally by Bruining et. al and partially solved in
[10], with same simpli-
fications as in the Example 2.1.1. The water, steam, nitrogen mass balance equation
19
and the energy balance equations are:
t
ϕρ
W
s
w
+
x
u
ρ
W
f
w
= q
g−→ a,w
, (2.10)
t
ϕρ
gw
s
g
+
x
u
ρ
gw
f
g
= q
g−→ a,w
, (2.11)
t
ϕρ
gn
s
g
+
x
u
ρ
gn
f
g
= 0, (2.12)
t
ρ
r
h
r
+
ϕ
(
ρ
W
h
w
s
w
+(
ρ
gw
h
gw
+
ρ
gn
h
gn
)s
g
)
+
x
u
ρ
W
h
w
f
w
+(
ρ
gw
h
gw
+
ρ
gn
h
gn
) f
g
= 0.
(2.13)
The quantities
ϕ
,
ρ
W
,
ρ
r
, s
w
, s
g
, f
w
and f
g
were defined in the previous example.
In the gaseous phase, the concentration of steam is
ρ
gw
(steam mass per unit gas
volume) and the concentration of nitrogen is
ρ
gn
; in the presence of liquid water ther-
modynamical considerations specify the dependence of these concentrations solely on
temperature, see Appendix A.
We consider an example illustrating how the accumulation and flux functions G
and F depend on the physical situation. We describe the gaseous physical situation,
where s
g
= 1 and s
w
= 0. We assume that nitrogen and steam behave as ideal
gases with densities denoted by
ρ
gN
and
ρ
gW
, see Appendix A. We also assume that
there are no volume effects due to mixing, namely the volumes of the components are
additive, i.e.,
ρ
gw
/
ρ
gW
(T)+
ρ
gn
/
ρ
gN
(T)=1. This fact allows to define the steam and
nitrogen gas composition, respectively, as:
ψ
gw
=
ρ
gw
/
ρ
gW
(T) and
ψ
gn
=
ρ
gn
/
ρ
gN
(T) , (2.14)
so the compositions are additive, i.e.,
ψ
gw
+
ψ
gn
= 1. We use
ψ
gw
as the independent
variable. Therefore, there are three unknowns to be determined: temperature T, gas
composition
ψ
gw
and speed u. The saturation is trivial, i.e., s
g
= 1 and s
w
= 0.
We rewrite equations (2.10)-(2.13) using Eq.
(2.14). Since s
w
= 0, f
w
= 0, from Eq.
(2.10) we obtain that q
g−→ a,w
vanishes, so the system is written as:
t
ϕ
M
W
p
at
R
ψ
gw
T
1
+
x
u
M
W
p
at
R
ψ
gw
T
1
= 0,
t
ϕ
M
N
p
at
R
(1
ψ
gw
)T
1
+
x
u
M
N
p
at
R
(1
ψ
gw
)T
1
= 0,
t
ϕ
ˆ
H
r
+
ψ
gw
M
W
p
at
RT
h
gW
+
ψ
gn
M
N
p
at
RT
h
gN
+
x
u
ψ
gw
M
W
h
gW
+
ψ
gn
M
N
h
gN
p
at
RT
= 0,
where
ˆ
H
r
=
ˆ
H
r
(T), h
gW
= h
gW
(T), h
gN
= h
gN
(T). The constants are described in
Appendix A. Notice that
(M
W
p
at
/R and M
N
p
at
/R are constants that cancel out of the
first two equations
In the single-phase gaseous situation, labelled by I, the spaces of primary and
trivial variables are
P
I
= {
ψ
gw
, T} and P
I
= {s
w
= 0} with
I
= {s
w
= 0,
ψ
gw
, T};
the Darcy speed u is an unknown (secondary variable). In the two-phase situation,
20
labelled by II, the spaces of primary and trivial variables are P
II
= {s
w
, T} and
P
II
= {
ψ
gw
(T)} with
II
= {s
w
,
ψ
gw
(T), T}; the Darcy speed u is an unknown (sec-
ondary variable). In this case the gas composition is given in terms of temperature by
Clausius-Clapeyron and Raoult’s laws, see
[44,68].
In the single-phase liquid situation, labelled by III, the spaces of primary and
trivial variables are
P
III
= {T} and P
III
= {s
w
= 1,
ψ
gw
(T)} with
III
= {s
w
=
1,
ψ
gw
(T), T}; the (secondary variable) Darcy speed is a constant. Even though there
is no gas in this physical situation, it is useful to define the gas composition by conti-
nuity from Eq
(2.14).
In each physical situation, it is necessary to calculate the solution for the primary
variables. In Figure 2.1, we show the physical situations in the primary state space
variable for I, II and III.In, Eqs.
(2.10)-( 2.13) represent a system with four equa-
tions and four variables: the temperature, the gas saturation, the gas composition
and the speed.
Steam region
Boiling region
Steam regionSteam region
Water region
Figure 2.1: State space, physical situations and their domains in Example 2.1.2. The
bold straight lines represent the nitrogen-free situations, that are: the watersituation
for s
w
= 1; the boiling situation for s
w
non-constant; the steam situation for s
w
= 0.
This Riemann problem is solved in Chapter 5.
2.1.1 Accumulations and fluxes for a physical situation
Generically, there is no unique choice of (G
j
, F
j
). As a simple example, consider G
j
=
(
G
j
1
, G
j
2
,···, G
j
n
(j)+1
) and F
j
=(F
j
1
, F
j
2
,···, F
j
n
(j)+1
) satisfying (2.2) for V
j
j
.A
21
permutation in the components of (G
j
, F
j
) yields the same solution of (2.2). However,
it is necessary to select a choice for the
(G
j
, F
j
) for each j ∈PS.
Property 2.1.1. We are interested in solving problems where there exists a linear map
E :
R
m+1
−→ R
n+1
that satisfies:
E
(Q(V)) = 0, ∀V, (2.15)
where Q
(V) is a (m + 1) ×1 vector and 0 is a (n + 1) ×1 vector. So the vector Q(V)
K(
E), where K(E) is the kernel of E.
For each chemical component, this map acts on
(2.1) by taking the equations for
all phases and adding them. The resulting equation
(2.2) represents the conservation
of mass of this component in all phases, so that the source terms add to zero.
Definition 2.1.7. We say that a linear map E :
R
m+1
−→ R
n+1
has maximal rank
associated to the system
(2.1),ormaximal rank, if it satisfies (2.15) and:
m
+ 1 = r ank(E)+di m(span{Q(V)}), (2.16)
where span
{Q} is the vector space generated by Q. The rank of E is denoted by n + 1.
We call the linear map E internal transfer matrix.
There exists a class of linear maps E :
R
m+1
−→ R
n+1
that satisfy (2.15) and (2.16)
and denote by E:
E = {E : R
m+1
−→ R
n+1
that satisfy (2.15) and (2.16)}. (2.17)
From
(2.16), for E ∈E, notice that K(E)=span{Q(V)}.
Definition of
(G
j
, F
j
)
Since this pair of functions is defined only in the physical situation j ∈PSand in
this physical situation we are interested only in the primary variables, it is necessary
to utilize
R
j
G
and R
j
F
defined in (2.6) and the relationship between the dependent
variable
V
j
and the corresponding primary variable V
j
. Let E ∈Ebe a linear map,
then for
V
j
and the corresponding primary variable V
j
, the pair (G
j
E
, F
j
E
) is:
G
j
E
(V
j
)=ER
j
G
(V
j
) and F
j
E
(V
j
)=ER
j
F
(V
j
). (2.18)
The number n
(j) represents the dimension of (G
j
E
, F
j
E
), i.e., the number of primary
variables to be determined.
Remark 2.1.4. Notice that the n
(j) is the same for any j ∈PS, so in what follows we
only write n to indicate that it does not depend on the physical situation.
From Lemma 2.1.1 and from the continuity of the linear map E we obtain:
22
Proposition 2.1.1. Consider two pairs (G
j
E
, F
j
E
) and (G
k
E
, F
k
E
) obtained from the same
linear map E from
(2.18), such that Π
jk
= Ø, then:
G
j
E
(V)=G
k
E
(V) and F
j
E
(V)=F
k
E
(V) V Π
jk
. (2.19)
Thus we can see that the cumulative and flux terms can be understood only as
functions of V for V belonging to some
P
j
.
Consider the balance system
(2.1) for V
j
j
, in a specified physical situation.
Applying the linear map E
∈Ein this system and using R
j
G
and R
j
F
defined in (2.6)
and (2.18), we obtain the following system of conservation laws (2.2) for which the
unknowns are the primary variables V
j
associated to V
j
that is:
t
G
j
E
+
x
uF
j
E
= 0. (2.20)
We would like the solution of the resulting conservation laws not to depend on
E
∈E, i.e., the solution of (2.2) should be the same for some E ∈E(at this stage, we
are not interested yet in the solution associated to initial and boundary conditions.
We are interested only in the form of the differential equation). To guarantee this
independence, we utilize the following Lemma:
Lemma 2.1.2. Let A and B be two vector spaces; E
1
: A −→ B and E
2
: A −→ B two
linear maps. If
K(E
1
)=K(E
2
), there exists an isomorphism I : B −→ B such that
E
1
= IE
2
.
Using Lemma 2.1.2, we obtain the following Proposition for E
∈E:
Proposition 2.1.2. Let E
1
, E
2
∈Ebe two different matrices. Then for (V
j
,u)
j
×R
for each j ∈PS, the corresponding (V
j
,u) ∈P
j
×R is the solution of
t
G
j
E
1
+
x
uF
j
E
1
= 0. (2.21)
if, only if, is solution of
t
G
j
E
2
+
x
uF
j
E
2
= 0, (2.22)
where V
j
is the primary variable associated to V
j
.
Proof: Since E
1
, E
2
∈E, then, by definition of E, we know that K(E
1
)=K(E
2
).
From Lemma 2.1.2 there exists an isomorphism
I : R
n+1
−→ R
n+1
such that E
1
= IE
2
.
Assume that
(V
j
,u) is a solution of (2.21). So using Def. 2.18 and E
1
= I
1
IE
1
=
I
1
E
2
, Eq. (2.21) can be written as:
I
1
E
2
t
R
j
G
+
x
u
R
j
F

= 0.
23
Since I
1
is also an isomorphism, K(I
1
)=0, so we obtain that:
E
2
t
R
j
G
+
x
u
R
j
F
= 0.
(2.23)
That is
(2.22). Since it is valid for E
1
and E
2
in E, the assertion follows.
Corollary 2.1.1. The solution of (2.2) does not depend on the linear map E ∈E.
Thus we can choose any E
∈Eand since the solution of conservation law does not
depend on E we rewrite
(G
j
E
, F
j
E
) as (G
j
, F
j
).
Example 2.1.3. For a conservation law in the form
(2.2), the transfer matrix E is any
non-singular matrix.
For the balance system (2.7)-(2.9
) and (2.10)-(2.13) the transfer matrices are :
E
=
110
001
and E
=
1100
0010
0001
.
2.1.2 Riemann problem
We use W
j
=(V
j
,u), for V
j
∈P
j
for some j ∈PSto represent the state variables,
where u is the secondary variable and V
j
are the primary state variables; similarly,
W =(V, u) for V∈ represents the state variables, u is the secondary variable
and
V are the primary together with the trivial state variables. We are interested in
the Riemann problem associated to
(2.1), that is the solution of these equations with
initial data
W
L
=(V
L
,u
L
) if x > 0,
W
R
=(V
R
,·) if x < 0.
(2.24)
Since the system
(2.1) has an infinite characteristic speed associated to u, only one of
the speeds u
L
or u
R
is given (in this case we have chosen u
L
). Later, it will be clear that
the other speed on the right hand side can be determined from the other equations in
the system.
The general solution of the Riemann problem associated to Eq.
(2.2) consists of a
sequence of elementary waves: rarefactions and shocks. These waves may be sepa-
rated by constant states. The solutions are studied in the sections that follow.
2.1.3 Wave Sequences and Riemann Solution
A Riemann solution is a sequence of elementary waves w
k
(shocks and rarefactions)
for k
= 1,2,···,m and constant states W
k
for k = 1,2,···,m.
W
L
≡W
0
w
1
−→ W
1
w
2
···
w
m
−→ W
m
≡W
R
. (2.25)
24
We represent any state by W. The wave w
k
has left and right states W
k1
and W
k
and speeds
ξ
k
<
ξ
+
k
in case of rarefaction waves and v =
ξ
k
=
ξ
+
k
in case of shock
waves. The left state of the first wave w
1
is (V
L
,u
L
) and the right state of w
m
is
(V
R
,u
R
), where u
R
needs to be found. In the Riemann solution it is necessary that
ξ
+
k
ξ
k+1
; this inequality is called geometrical compatibility. When
ξ
+
k
<
ξ
k+1
there
is a separating constant state W
k+1
between w
k
and w
k+1
; in this sequence the wave
w
k
is indicated by .If
ξ
+
k
=
ξ
k+1
there is no actual constant state in physical space,
so the wave w
k
is a composite with w
k+1
; it is indicated by .
Different physical situations are separated by shocks respecting the geometrical
compatibility. When it is useful to emphasize the waves in the sequence
(2.25) rather
than the states, we use the notation:
w
1
w
2
··· w
m
, (2.26)
where for each w
k
, represents for the case
ξ
+
k
<
ξ
k+1
and for the case
ξ
+
k
=
ξ
k+1
.
2.2 Characteristic speeds
Since the cumulative and flux terms are not smooth between physical situations, we
calculate the characteristic speeds only within the physical situation. The system
of conservation reduces to the conservation form
(2.2), thus it is possible to find the
characteristic speeds, and then the rarefaction waves. These waves are important to
solve the Riemann problem; the characteristic speeds are also important to propose
adequate numerical methods to solve Eq. (2.1) and condition for numerical stability,
see
[40].
For a smooth wave we differentiate all equations in (2.2) with respect to their
variables, obtaining a system:
B
t
V
u
+ A
x
V
u
= 0; (2.27)
the matrices B and A are the derivatives of G
(V) and uF(V) with respect to W =
(
V,u), where V = V
j
∈P
j
are the primary variables for some j ∈PSand u is the
speed. Since G
(V) does not depend on u, the matrix G/W has a zero column, ie.
G
W
=
G
1
V
1
G
1
V
2
···
G
1
V
n
0
G
2
V
1
G
2
V
2
···
G
2
V
n
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
G
n
V
1
G
n
V
2
···
G
n
V
n
0
G
n+1
V
1
G
n+1
V
2
···
G
n+1
V
n
0
=
G
V
0
:
= B, (2.28)
25
and
uF
W
=
u
F
1
V
1
u
F
1
V
2
··· u
F
1
V
n
F
1
u
F
2
V
1
u
F
2
V
2
··· u
F
2
V
n
F
2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
u
F
n
V
1
u
F
n
V
2
··· u
F
n
V
n
F
n
u
F
n+1
V
1
u
F
n+1
V
2
··· u
F
n+1
V
n
F
n+1
=
u
F
V
F
:
= A, (2.29)
where G
/V and uF/V represent the n first columns of B and A, 0 and F represents
the
(n + 1)-th column of B and A.
Remark 2.2.1. Let V
∈P
j
.IfV
is private point, the derivatives G/V and F/V
are well defined at V
, because if the point is an internal point the cumulative terms G
and flux terms F are
C
2
. On the other hand, if V
is a boundary point the derivatives
are defined laterally.
However, if V
is a frontier point, i.e., there exists another k ∈PS, with k = j,
such that V
∈P
k
,atV
the derivatives are not well defined, because the derivatives
can be different in neighboring physical situations. Since by Rem. 2.1.3 each physical
situation is closed, and we are interested in the rarefaction waves in a determined
physical situation, we define the derivatives laterally in each
P
j
.
The eigenvalues
λ
and right eigenvectors
r =(r
1
,r
2
,···,r
n+1
)
T
with n + 1 compo-
nents of the following system are the rarefaction wave speeds and the characteristic
directions, respectively:
A
r =
λ
B
r. (2.30)
Hereafter, the word “eigenvectors” means “right eigenvectors”. Similarly the left
eigenvector
=(
1
,
2
,···,
n+1
) satisfies:
A =
λ
B or A
T
T
=
λ
B
T
.
To find
λ
, we need to solve the characteristic equation
det
(A
λ
B)=0. (2.31)
Whenever necessary, we write the right and left eigenvectors without the upper arrow,
i.e., r
=
r and =
.
Remark 2.2.2. We utilize an index p to indicate a particular characteristic speed fam-
ily. The eigenvalue, right eigenvector and left eigenvector of the p
family are denoted
by
λ
p
, r
p
=(r
p
1
,r
p
2
,···,r
p
n
,r
p
n
+1
) and
p
=(
p
1
,
p
2
,···,
p
n
,
p
n
+1
).
We define the set of indices n as
C:
C = {1,2,···,n + 1}. (2.32)
26
Definition 2.2.1. The integral curve of the p-family is the solution of :
dV
d
ξ
,
du
d
ξ
= r
p
, i.e,
dV
i
d
ξ
= r
p
i
for all i ∈C and
du
d
ξ
= r
p
n
+1
. (2.33)
When
ξ
increases satisfying:
ξ
=
λ
p
(V(
ξ
),u(
ξ
)), (2.34)
the integral curve defines a rarefaction curve. If in addition to
(2.34), the variable
ξ
satisfies x =
ξ
t the integral curve defines rarefaction waves in the (x, t) plane.
Lemma 2.2.1. Assume that u
= 0 and that F
k
(V) = 0 for some fixed k ∈Cfor all V.
The eigenvalue, right and left eigenvectors for the system
(2.27), with B and A given
by Eqs.
(2.28) and (2.29) have the form:
λ
= u
ϑ
(V), (2.35)
r
=(g
1
(V), g
2
(V) ··· , g
n
(V), ug
n+1
(V)) and =(
1
(V),
2
(V), ··· ,
n+1
(V)),
(2.36)
where
ϑ
, g
i
and
i
(for all i ∈C) are functions that depend only on V. Moreover, there
are at most n eigenvalues and eigenvectors associated to this system of n
+ 1 equations.
Proof: The eigenvalues
λ
of (2.27) are the roots of det(A
λ
B)=0, for B and A
given by Eq.
(2.28) and (2.29). This characteristic equation is:
det
(A
λ
B)=det
u
F
V
(V)
λ
G
V
(V)
F
(V)
= 0. (2.37)
Since u
= 0, we divide the first n columns by u and define
ϑ
=
λ
/u, then after some
calculations Eq.
(2.37) reduces to:
det
F
V
(V)
ϑ
G
V
(V)
F
(V)
= 0. (2.38)
Since Eq
(2.38) depends on V only, the eigenvalues
ϑ
are functions of V, i.e.,
ϑ
=
ϑ
(V),
they do not depend on u. Using that
λ
= u
ϑ
, the eigenvalues have the form (2.35).
The eigenvectors are solutions of
(A
λ
B)r = 0. Writing r =(r
1
,···,r
n+1
)
T
,we
obtain:
u
F
V
(V)
ϑ
(V)
G
V
(V)
F
(V)
r
= 0. (2.39)
Since there is an index k such that F
k
(V) = 0, then we can write r
n+1
as:
r
n+1
= u
1
F
k
(V)
n
l=1
F
k
V
l
(V)
ϑ
(V)
G
k
V
l
(V)
r
l
, (2.40)
Substituting r
n+1
given by Eq. (2.40) into Eq. (2.39), we obtain a linear system in
the unknowns r
j
for j = 1,2,··· ,n, where we can cancel u, showing that the r
l
for
l
= 1,2,···,n depend on V only. So (2.36. a) follows from (2.40).
27
For =(
1
,
2
,···,
n+1
), we solve:
·
u
F
V
(V)
ϑ
(V)
G
V
(V)
F
(V)
= 0,
then
is solution of the following system:
u
F
1
V
l
ϑ
G
1
V
l
1
+ ···+ u
F
n+1
V
l
ϑ
G
n+1
V
l
n+1
= 0 for l = 1,··· ,n (2.41)
F
1
1
+ F
2
2
+ ···+ F
n+1
n+1
= 0. (2.42)
Since u = 0, we divide Eqs. (2.41) by u, then we obtain a system with coefficients that
depend only on the variables V, leading to
(2.36.b).
Remark 2.2.3. The hypothesis that there exists an index k such that F
k
(V) = 0 for all
V is sufficient, but it is not necessary. However in the most of problems this hypothesis
is satisfied.
Proposition 2.2.1. Assume that locally the eigenvector r associated to a certain family
forms a local vector field. Then we calculate the primary variables V on the rarefaction
waves in the
(x, t) plane independently of u, i.e., first we obtain the primary variables
in the classical way, i.e., solving
(2.33.b), and then we calculate the secondary variable
u in terms of the primary variables from:
u
= u
ex p(
γ
(
ξ
)),
γ
(
ξ
)=
ξ
ξ
g
n+1
(V(
η
))d
η
, (2.43)
where
ξ
= x/t,
ξ
=
λ
(W
) and u
is the “first” value of u on the rarefaction wave,
i.e., u
= u
for
ξ
=
ξ
.
Proof: Using
(2.33) and (2.34) and since g
i
(V)=r
i
for i = 1,2,···,n we obtain V
independently of u by solving the system of differential equations:
dV
d
ξ
=
dV
1
d
ξ
,
dV
2
d
ξ
,···,
dV
n
d
ξ
=(g
1
, g
2
,···, g
n
), with
ξ
=
λ
(W
).
After obtaining V
(
ξ
), we use the expression for the last component of r in (2.36.a)
to solve du/d
ξ
= ug
n+1
, yielding (2.43). From Lemma 2.2.1, we know that
λ
has the
form u
ϑ
(V), so using Eqs. (2.34) and (2.43), we obtain
ξ
implicitly as:
ξ
= u
ϑ
(V(
ξ
))ex p(
γ
(
ξ
)). (2.44)
Since
ξ
depends only on u
and V on the integral curves, the results follows.
Definition 2.2.2. Since only the first n coordinates of eigenvectors are important to
obtain the integral curves in the space V (i.e. independently of u), it is useful to define
for any p-family the structures, see Rem. 2.2.2:
ˆ
r
p
=(r
p
1
,r
p
2
,···,r
p
n
) and
ˆ
p
=(
p
1
,
p
2
,···,
p
n
), (2.45)
and
ˆ
λ
p,
(V) :=
λ
p
(W)/u
and
ˆ
λ
p,+
(V) :=
λ
p
(W)/u
+
. (2.46)
28
Corollary 2.2.1. Assume that u
= 0, so we perform the change of variables:
ξ
=
x
t
−→
ˆ
ξ
=
x
u
t
, (2.47)
The system
(2.2) can be written in the space of variables (x, T)=(x, u
t).
Proof: In V space, one can prove that the eigenvalues and eigenvectors have the
form
λ
(V), with r given in Eq. (2.36. a). From Eqs. (2.34) and (2.44), it follows that
ˆ
ξ
and
ˆ
ξ
satisfy:
ˆ
ξ
=
ϑ
(V
) and
ˆ
ξ
=
ϑ
(V(
ˆ
ξ
))ex p(
γ
(V(
ξ
))). (2.48)
Thus in the
(x, T) plane, the variables V and
ˆ
ξ
do not depend on u, so the rarefaction
curve does not depend on u. The speed keeps the form
(2.43).
Corollary 2.2.2. If r
n+1
0 in a region, then from Prop. 2.2.1 u is constant along the
corresponding rarefaction curve.
2.3 Shock waves
Typically, discontinuities or shocks appear in solutions of Riemann (or Cauchy) prob-
lems of non-linear hyperbolic or hyperbolic-elliptic equations (see
[60]). The Rankine-
Hugoniot (RH) condition for shocks needs to be satisfied as it expresses conservation
of mass. For a given state W
, the set of states W
+
that satisfies the RH condition is
called Rankine-Hugoniot locus (RH locus) of W
, which is denoted by RH (W
).
2.3.1 Shocks and the Rankine-Hugoniot condition
The main feature of the class of equations (2.1) is the existence of separate regions
where the system of balance equations (2.1
) reduces to different systems of conserva-
tion equations
(2.2). To obtain the complete Riemann problem solution it is necessary
to link these regions. Because often the term Q gives rise to fast changes between
regions, we consider shocks linking these different regions.
We apply a linear map E given in Def. 2.1.7 on the system
(2.1). Since E[Q]=0,
we obtain a system in conservative form:
t
E
G(V)+
x
(
uEF(V)
)
= 0, (2.49)
where
V = V
j
j
for some j ∈PS.
Remark 2.3.1. We should use a single E linking neighboring physical situations. No-
tice that it is possible to use a single E for the complete system yielding the simplest
global formulation. However, we have proved that the weak solution does not depend
on E, so we can use different linear maps E in the interior of each physical situation,
see Proposition 2.1.2.
29
The RH condition applied to Eq. (2.49) is written as:
v
s
E
G(V
+
) EG(V
)
= u
+
EF(V
+
) u
EF(V
), (2.50)
where
(V
,u
) is the state on the left of the shock and (V
+
,u
+
) is the state on the
right of the shock; v
s
is the shock speed. We specify the left conditions for the variables
V and u, but as we will see, the right conditions are specified only for one component
of
V
+
. The speed u
+
is always obtained from the RH condition (2.50).
The define the cumulative mass transfer in the shocks can be defined as:
[Q(V)] := u
+
F(V
+
) u
F(V
) v
s
G(V
+
) −G(V
)
. (2.51)
Once the Riemann solution is computed,
(2.51) calculates the rate of transfer of each
chemical component in a shock.
We consider shocks of two different types: shocks within regions in a certain phys-
ical situation and shocks between different physical situations. In both cases V
and
V
+
are the primary variables associated to V
and V
+
, respectively.
Shock within a physical situation
Since
V
+
,V
j
for some j ∈PS, using the definition of R
j
G
and R
j
F
given in Eq.
(2.6) and the relationship (2.18) , the RH condition (2.50) can be written as:
v
s
G
j
(V
+
) G
j
(V
)
= u
+
F
j
(V
+
) u
F
j
(V
). (2.52)
Shock between distinct physical situations
Since
V
+
j
and V
k
for some pair j, k ∈PSwith k = j, using the definition of
R
j
G
, R
k
G
R
j
F
and R
k
F
given in Eq. (2.6) and the relationship (2.18), the RH condition
(2.50) can be written as:
v
s
G
j
(V
+
) G
k
(V
)
= u
+
F
j
(V
+
) u
F
k
(V
). (2.53)
In the following sections, we will always indicate functions of the shock left state
by the superscript
() and functions of the right state of the shock by (+). This
notation encompasses both
(2.52) and (2.53).
2.3.2 The Rankine-Hugoniot locus and shock waves
Notice that in the RH condition (2.52) or (2.53) for each fixed state (), there is a
degree of freedom in the variables V, that is a 1-dimensional structure, generically.
This structure is called the RH locus.
30
Writing Eq. (2.53) as a linear homogeneous system, we obtain:
[
G
1
]
F
+
1
F
1
[
G
2
]
F
+
2
F
2
.
.
.
.
.
.
.
.
.
[
G
n
]
F
+
n
F
n
[
G
n+1
]
F
+
n+1
F
n+1
v
s
u
+
u
= 0, or M
v
s
u
+
u
= 0, (2.54)
where
[G
i
] for i ∈Cis given by:
[
G
i
]
=
G
j
i
(V
+
) G
k
i
(V
), F
+
i
= F
j
i
(V
+
), F
i
= F
k
i
(V
), (2.55)
The 3
×3 minors of the matrix M are denoted by M
pqs
:
M
pqs
=
G
p
F
+
p
F
p
G
q
F
+
q
F
q
[
G
s
]
F
+
s
F
s
for all distinct p, q and s in
C. (2.56)
The homogeneous linear system
(2.54) has a non-trivial solution, if only if
H
pqs
= det
M
pqs
= 0 for all distinct p, q and s in C. (2.57)
Definition 2.3.1. The RH locus is given by implicit expressions in the primary vari-
ables. For
(V
,u
) fixed in some physical situation, the RH locus in the projected
space V consists of the V
+
that satisfy det(M
pqs
)=0 for all distinct p, q and s in C.
Notice that V
+
can lie in the same or in a distinct physical situation relatively to V
.
We denote the RH locus of a state V
as RH(V
).
Notice that there are
C
3
n
+1
equations satisfying (2.57); however, under certain
hypotheses, the following Lemma guarantees that we only need to solve a set of
(n1)
equations of type (2.57).
Since we want to characterize the solution of
(2.57), it is useful to define for each
i
∈C, the vector
D
i
≡D
i
(V
;V
+
)
[G]
i
,F
+
i
, F
i
. (2.58)
Using the notation
(2.58) we obtain the simple lemma:
Lemma 2.3.1. Let V
be fixed. Assume that there exists two linearly independent
vectors
D
l
and D
m
for l,m ∈Cin a neighborhood B
j
δ
(V
+
). If there are (n 1) dis-
tinct minors
M
lms
for s ∈C{l, m} such that det(M
lms
)=0, then it follows that
det
(M
pqs
)=0 for all distinct p, q and s in C.
Proof: There are
(n 1) indices s ∈C−{l, m} such that det(M
lms
)=0.
Since
D
l
and D
m
are L.I., it follows that for all s ∈Cthe vector D
s
is a linear
combination of
D
l
and D
m
. Since any minor M
pqs
with distinct p, q and s in C has the
form:
M
pqs
=
D
p
D
q
D
s
,
it follows that det
(M
pqs
)=0 and the lemma is proved.
31
Definition 2.3.2. Consider V
∈P
j
and V
+
∈P
k
. Assume that V ∈P
k
belongs to a
1-dimensional locus where the vectors
D
p
and D
q
are linearly dependent (LD) for some
p
= q ∈C. We call this locus in a linear degeneracy and denote it by X
pq
(V
;V
+
).
Example 2.3.1. In Section 4.5, we show that in the single-phase gaseous situation (see
Example 2.1.2), the accumulations and fluxes for steam and nitrogen are:
G
1
=
ϕθ
W
ψ
gw
T
, F
1
=
ϕθ
W
ψ
gw
T
, G
2
=
ϕθ
N
ψ
gn
T
and F
2
=
ϕθ
N
ψ
gn
T
,
where
ψ
gn
= 1
ψ
gw
. The constans
θ
W
,
θ
N
are given by by Eq. (4.24). The terms D
1
and D
2
are:
D
1
=
ϕθ
W
ψ
+
gw
T
+
ψ
gw
T
ϕθ
W
ψ
+
gw
T
+
ϕθ
W
ψ
gw
T
D
2
=
ϕθ
N
ψ
+
gn
T
+
ψ
gn
T
ϕθ
N
ψ
+
gn
T
+
ϕθ
N
ψ
gn
T
.
Let
(T
,
ψ
gw
) be fixed. It is easy to see that in the plane (T,
ψ
gw
) the straight line
(T,
ψ
gw
=
ψ
gw
) is the linear degeneracy X
12
(V
). Of course, this straight line belongs
to the RH locus for
().
Remark 2.3.2. If a linear degeneracy occurs a more detailed analysis is required.
Corollary 2.3.1. Given V
∈P
j
for some j ∈PS, assume that for V
+
∈RH(V
)
and in a neighborhood B
j
δ
(V
+
) there exists at least a pair of linearly independent (LI)
vectors
D
p
and D
q
with p,q ∈Cof the form (2.58). Assume also that for all l = m ∈C
and V
+
∈B
j
δ
(V
+
) X
lm
(V
;V
+
) the vectors D
l
and D
m
are linearly independent.
Then the RH locus is a 1-dimensional structure and it is obtained as the solution of
(n 1) equations:
det
D
p
D
q
D
r
= 0, r ∈C−{p,q}. (2.59)
Moreover, X
lm
(V
;V
+
) for l = m ∈Cbelongs to the RH locus.
Definition 2.3.3. The minimum set formed by the triplet of indices that define implic-
itly the RH locus are called RH indices and they are denoted by
C
RH
. There are n 1
triplets, numbered from
1 to
−−→
n 1. There is no a unique choice for C
RH
.
So using Def.
(2.57), the RH locus is:
H = 0 for all l ∈{
1,
2,···,
−−→
n 1}∈C
RH
, (2.60)
where each number l corresponds to a triplet in
C
RH
.
32
Let us assume that RH (V
) was found. Now we want to determine u
+
. Let
p,q
∈C. Given V
k
for some k ∈PS,ifD
p
and D
q
are LI in a neighborhood
B
j
δ
(V
+
), then:
det
[G
p
] F
+
p
[G
q
] F
+
q
= 0. (2.61)
The shock and speeds are obtained by solving the following system on the RH locus:
[G
p
] F
+
p
[G
q
] F
+
q

v
s
u
+
=
u
F
p
u
F
q
. (2.62)
In this case v
s
and u
+
can be written as function of u
and V
as:
v
s
= u
F
+
p
F
q
F
+
q
F
p
F
+
q
[G
p
] F
+
p
[G
q
]
and u
+
= u
F
q
[G
p
] F
p
[G
q
]
F
+
q
[G
p
] F
+
p
[G
q
]
. (2.63)
Of course, we could have chosen to solve v
s
and u
in terms of u
+
.
Remark 2.3.3. In the definitions that follow, all wave structures can be obtained
solely in the space of primary variables V. Using u
+
:= u
+
(V
,u
;V
+
) and v
s
:=
v
s
(V
,u
;V
+
), we define:
u
+
u
=
χ
(V
;V
+
),
ˆ
v
(V
;V
+
) :=
v
s
u
and
ˆ
v
+
(V
;V
+
) :=
v
s
u
χ
(V
;V
+
)
. (2.64)
Let us index the rarefaction curves in the Riemann solution by l for l
= 1,2,···,
ρ
1
;
the shocks are indexed by m for m
= 1,2,···
ρ
2
. From the Sections 2.2 and 2.3.2,we
have proved the following:
Proposition 2.3.1. Let u
L
be positive. The primary variables V in the shock and
rarefaction curves do not depend on the left speed u
L
> 0. If a sequence of waves and
states solve the Riemann problem in the primary variables, for a given u
L
> 0 (or
u
R
> 0), then it is also a solution for any other u
L
> 0 (or u
R
> 0):
(V
L
,u
L
) if x < 0
(V
R
,·) if x > 0.
(2.65)
Moreover, assume that for each m there are p, q
∈Csuch that the following inequality
is satisfied: F
+
q,m
[G
p,m
] F
+
p,m
[G
q,m
] = 0. Then u
R
is given by:
u
R
= u
L
ρ
1
l=1
ex p
ξ
+,l
ξ
,l
g
l
n
+1
(V(
η
))d
η
ρ
2
m=1
F
q,m
[G
p,m
] F
p,m
[G
q,m
]
F
+
q,m
[G
p,m
] F
+
p,m
[G
q,m
]
. (2.66)
where g
l
n
+1
is the (n + 1)-th component of eigenvector r,
ξ
,l
and
ξ
+,l
are the first and
the last values of
ξ
l
associated to the l-th rarefaction wave.
33
In the Proposition above, G
j,m
and F
j,m
represent the j-th components ( j ∈C) of G
and F on the m-th shock wave. Similarly ex p
ξ
l
ξ
0,l
g
n+1
(V(
η
))d
η
is computed along
the l-rarefaction curve, see
(2.43).
In Proposition 2.3.1, instead of u
L
we could have prescribed the right speed u
R
and
obtain u
L
and the other speeds as function of V and u
R
.
Proposition 2.3.2. Assume that u
L
= 0. If the speed u
L
in the initial data is modified
while
V
L
and V
R
are kept fixed, the speed u
R
as well as the Riemann solutions are
rescaled in the
(x, t) plane, while the wave sequences and the values of V are kept
unaltered.
Proof: The rarefaction case was proved in Corollary 2.2.1. In the shock case,
performing the change of variable
(2.47) with u
= u
L
and using (2.52):
v
s
u
L
(G
+
G
)=
u
+
u
L
F
+
u
u
L
F
. (2.67)
Defining U :
= u/u
L
and
ˆ
v
L
= v
s
/u
L
, we have:
(1) If there exists a rarefaction before u
, from Corollary 2.2.1, we see that u
=
u
L
T (V), where T (V) := ex p(
γ
(V(
ξ
))) depends only on V.
(2) If there is no rarefaction with u
= u
L
, we conclude that (2.67) can be rewritten
as:
ˆ
v
L
(G
+
G
)=U
+
(V)F
+
U
(V)F
, (2.68)
where U
(V)=T (V) if (1) is satisfied or U
(V)=1 if (2) is satisfied. Since U
+
(V)
depends only on V, using Corollary 2.2.1 we obtain the result.
This Proposition leads to an important result:
Corollary 2.3.2. Assume the same hypotheses of Proposition 2.3.2. Then the Riemann
solution can be obtained in each physical situation first in the primary variables V.
The solution for the speed can be obtained in terms of
V and u
L
by Eq. (2.66).The
same is true for rarefaction and shock speeds.
We can now prove a generalization of the Triple Shock Rule
[25]:
Proposition 2.3.3. (Triple Shock Rule.) Consider three states in different physical
situations with two speeds namely
V
M
,u
+
,
(
V
+
,u
+
)
and
(
V
,u
)
. Assume that
V
∈RH(V
+
), V
M
∈RH
(
V
)
and V
+
∈RH(V
M
), with shock speeds v
+,
, v
,M
and v
M,+
. Then, either:
(1) v
+,
= v
M,
= v
+,M
;or
(2) G
(V
+
) G(V
) and G(V
+
) G(V
M
) are LD.
Proof: The RH conditions can be written as:
v
+,
(G(V
) G(V
+
)) = u
F(V
) u
+
F(V
+
), (2.69)
v
,M
(G(V
M
) G(V
)) = u
+
F(V
M
) u
F(V
), (2.70)
34
v
M,+
(G(V
+
) G(V
M
)) = u
+
F(V
+
) u
+
F(V
M
). (2.71)
Add
(2.71), (2.70) and (2.69). We obtain:
(v
M,+
v
,M
)
G
(V
+
) G(V
M
)
+(v
,M
v
+,
)
G
(V
+
) G(V
)
= 0. (2.72)
Thus
(1) or (2) are satisfied.
For future use we define := (V
,u
;V
M
,u
M
;V
,u
;V
+
,u
+
) as:
:=(v
M,
v
,+
)
G
(V
+
) G(V
)
+(v
,M
v
M,
)
G
(V
M
) G(V
)
; (2.73)
The limitation of the triple shock rule is that, in several applications we do not
know the intermediate speed. In such cases we need the following result:
Proposition 2.3.4. (Quadruple Shock Rule I.) Consider two physical situations, with
three primary points and three speeds, determining four states, namely
(V
,u
) in
k
,
(V
+
,u
+
) and (V
+
,u
) in
j
and (V
M
,u
M
) either in
j
or
k
. Assume that there are
shocks between the following pairs of states:
(i)
(V
,u
) and (V
+
,u
+
) with speed v
,+
,
(ii)
(V
,u
) and (V
M
,u
M
) with speed v
,M
,
(iii)
(V
M
,u
M
) and (V
+
,u
) with speed v
M,
.
We conclude that:
R1) If
= 0 and at least two shock speeds coincide, i.e., one of the following shock
speed equalities is satisfied:
either v
,+
= v
,M
or v
,+
= v
M,
or v
,M
= v
M,
. (2.74)
Then:
(1) u
= u
+
;
(2) all three shock speeds are equal:
v
,+
= v
,M
= v
M,
. (2.75)
R2) On the other hand, if
= 0, then u
= u
+
.
Proof: Assume first G
(V
+
) G(V
) and G(V
M
) G(V
) are LI. Disregarding
the indices that indicate the physical situations in the cumulative and flux terms, the
RH condition for
(V
,u
)-(V
+
,u
+
), (V
,u
)-(V
M
,u
M
) and (V
M
,u
M
)-(V
+
,u
) can be
written, respectively as:
v
,+
(G(V
+
) G(V
)) = u
+
F(V
+
) u
F(V
), (2.76)
v
,M
(G(V
M
) G(V
)) = u
M
F(V
M
) u
F(V
), (2.77)
v
M,
(G(V
+
) G(V
M
)) = u
F(V
+
) u
M
F(V
M
). (2.78)
35
Assume now that Eq. (2.74. a) is satisfied. Substituting v
,+
= v
,M
= v in Eqs.
(2.76
) and (2.77) and subtracting the resulting equations, we obtain:
v
(G(V
+
) G(V
M
)) = u
+
F(V
+
) u
M
F(V
M
), (2.79)
From Eq.
(2.63), notice that the shock speeds depend solely of the primary vari-
ables and speed on the left state and on the primary variable of the right state. From
Eqs.
(2.78) and (2.79), we can see that the left and right states as well as left speed
are the same, so u
= u
+
. Eq. (2.75) is satisfied also because the shock speed depends
just on the initial and final states.
To prove R2, we subtract
(2.76) from the sum of (2.77) with (2.78), obtaining:
=(u
+
u
)F(V
+
).
If
= 0, then u
= u
+
and the result follows.
The other cases are proved similarly.
Remark 2.3.4. The equality = 0 is a particular case of the Triple Shock Rule.
Generically, we do not know if V
+
∈RH(V
)
RH(V
M
). Therefore, we rewrite
Prop. 2.3.4 in the more practical form:
Proposition 2.3.5. (Quadruple Shock Rule II.) Consider two physical situations
j
and
k
for k, j ∈PS, with four primary points and four speeds, determining four
states, namely
(V
,u
) in
k
, (V
+
,u
+
) in
j
, (V
M
,u
M
) and (V
,u
) either in
j
or
k
. Assume that there are shocks between the following pairs of states:
(i)
(V
,u
) and (V
+
,u
+
) with speed v
,+
,
(ii)
(V
,u
) and (V
M
,u
M
) with speed v
,M
,
(iii)
(V
M
,u
M
) and (V
,u
) with speed v
M,
,
Assume also that the following conditions are satisfied:
(I) G
(V
+
) G(V
) and G(V
) G(V
M
) are LI;
(II) two shock speeds coincide, i.e., at least one of the following shock speed equali-
ties is satisfied:
either v
,+
= v
,M
or v
,+
= v
M,
or v
,M
= v
M,
. (2.80)
(III) V
+
and V
have a coordinate V
i
with coinciding values;
(IV) The curve
RH(V
M
) does not possess two different points with the same coordi-
nate V
i
having coinciding values.
Then
(1) V
=V
+
;
(2) u
= u
+
;
(3) all three shock speeds are equal:
v
,+
= v
,M
= v
M,
. (2.81)
36
Proof: Disregarding the indices that indicate the physical situations in the accu-
mulation and flux terms, the RH conditions for
(V
,u
)-(V
+
,u
+
), (V
,u
)-(V
M
,u
M
)
and (V
M
,u
M
)-(V
,u
) are written, respectively as:
v
,+
(G(V
+
) G(V
)) = u
+
F(V
+
) u
F(V
), (2.82)
v
,M
(G(V
M
) G(V
)) = u
M
F(V
M
) u
F(V
), (2.83)
v
M,
(G(V
) G(V
M
)) = u
F(V
) u
M
F(V
M
). (2.84)
Assume that now Eq.
(2.80.a) is satisfied. Substituting v
,+
= v
,M
= v in Eqs.
(2.82
) and (2.83) and subtracting the resulting equations, we obtain:
v
(G(V
+
) G(V
M
)) = u
+
F(V
+
) u
M
F(V
M
). (2.85)
Notice that Eqs.
(2.85) and (2.84) define implicitly the RH locus from (V
M
,u
M
). Since
the RH locus in the variable V depends solely on V
M
and the accumulation and flux
functions, the RH locus defined by Eqs.
(2.85) and (2.84) coincide. Because (III) and
(IV) are satisfied, it follows that V
= V
+
.
Now from Eqs.
(2.64), notice that the and shock speeds depend solely on the pri-
mary variables and speed of the left state and on the primary variable of the right
state. From Eqs.
(2.84) and (2.85), we can see that the left and the right states are
the same for each expression and define the same RH locus, so u
= u
+
and Eq. (2.81)
is satisfied.
The other cases are proved similarly.
2.3.3 Extension of the Bethe-Wendroff theorem
There are loci where the solutions change topology such as: secondary bifurcations,
coincidences, double contacts, inflections, hystereses and interior boundary contacts.
So the Riemann problem cannot be solved by the classical construction
[43]. The
Bethe-Wendroff theorem is an important tool to characterize such loci.
We prove an extension of the important Bethe-Wendroff theorem for shocks be-
tween regions in different situations. We recall that the left eigenvector
does not
depend on u. We recall the notation
[G]=(G
+
G
), where G
+
= G(V
+
) and
G
= G(V
).
Proposition 2.3.6. Let v
s
(W
+
;W
) be the shock speed between different physical sit-
uations. Assume that
p
(V
+
) · [G] = 0. Then v
s
has a critical point at W
+
and so
ˆ
v
+
(V
,V
+
) has a critical point at V
+
, if and only if:
ˆ
v
+
(V
;V
+
)=
λ
p,+
(V
+
) for the family p, (2.86)
where
λ
p,+
(V) is given by Eq. ( 2.46) and
ˆ
v
+
(V
;V
+
) is given by (2.64.c).
Proof: Assuming that the RH curve can be parametrized by
ζ
in a neighborhood
of
(V
+
,u
+
), we can write the RH condition as:
v
s
(G(V(
ζ
)) G
)=u(
ζ
)F(V(
ζ
)) u
F
, (2.87)
37
where v
s
:= v
s
(
ζ
). Differentiating (2.87) with respect to
ζ
we obtain:
dv
s
d
ζ
(G(V(
ζ
)) G
)+v
s
G(V(
ζ
))
W
dW
d
ζ
=
(
u(
ζ
)F(V(
ζ
))
)
W
dW
d
ζ
, (2.88)
Setting
ζ
+
such that (V(
ζ
+
),u(
ζ
+
)) = W(
ζ
+
)=W
+
=(V
+
,u
+
), Eq. (2.88) yields:
[G]
dv
s
d
ζ
+ v
s
G
W
dW
d
ζ
=
(
uF
)
W
dW
d
ζ
, (2.89)
where W
=(V, u). Assume first that (2.86) is satisfied. Notice that if
ˆ
v
+
(V
,V
+
)=
ˆ
λ
+
, then for (V
+
,u = u
+
),wehave
λ
= u
+
ˆ
λ
+
and v
s
(V
,u
;V
+
)=
λ
(V
+
,u
+
) (we
dropped the family index p). Substituting them in
(2.89) we obtain at (+) = (V
+
,u
+
)
the expression (2.89) with v
s
:= u
λ
/u. Let the left eigenvector associated to
ˆ
λ
+
;
taking the inner product of
(2.89) at (V
+
,u
+
) by , we obtain:
· [G]
dv
s
d
ζ
+ ·
u
λ
u
G
W
(
uF
)
W
dW
d
ζ
= 0. (2.90)
Since
is an eigenvector associated to
λ
, the second term of (2.90) is zero and it follows
that:
·[G]
dv
s
d
ζ
= 0.
Since by hypothesis
· [G] = 0, we obtain that dv
s
/d
ζ
= 0 and the shock speed is
critical.
On other hand, assume that v
s
has a critical point, so dv
s
/d
ζ
= 0 and Eq. (2.89)
reduces to:
v
s
G
W
dW
d
ζ
=
(
uF
)
W
dW
d
ζ
, or
(
uF
)
W
v
s
G
W
dW
d
ζ
= 0. (2.91)
Eq.
(2.91) has a solution if, only if,
v
s
(V
,u
;V
+
)=
λ
(V
+
,u
+
) so
ˆ
v
+
(V
;V
+
)=
ˆ
λ
+
(V
+
).
Following Oleinik [53] and Liu [47,48], we can obtain a useful result for equations
in the form
(2.2):
Proposition 2.3.7. Let
ˆ
v
+
(V
+
;V
) be the shock speed. Assume that
+
·(G
+
G
) =
0 for , r and
λ
belong to the same family p. Assume that
ˆ
v
+
has a critical point at V
+
.
Then the following relations are true:
i) If
ˆ
v
+
has a minimum:
(
+
·
W
G
+
r
+
)(
λ
·r
+
)
+
·[G]
>
0. (2.92)
38
ii) If
ˆ
v
+
has a maximum:
(
+
·
W
G
+
r
+
)(
λ
·r
+
)
+
·[G]
<
0, (2.93)
where
W
G
+
:=
(
G/W
)
(
V
+
),
+
:= (V
+
) and r := r(V
+
,u
+
). Moreover the in-
equalities
(2.92) and (2.93) do not depend on the speed u.
Proof: Assume that the shock and rarefaction curves can be parametrized by
ζ
in a neighborhood of (V
+
,u
+
) (if the curves exist, these parametrizations exists in a
small neighborhood of
(V
+
,u
+
)).
Differentiating Eq.
(2.88) again we obtain:
d
2
v
s
d
2
ζ
(G(V(
ζ
)) G
)+2
dv
s
d
ζ
G(V(
ζ
))
W
dW
d
ζ
+ v
s
2
G(V(
ζ
))
2
W
dW
d
ζ
,
dW
d
ζ
+
+
v
s
G(V(
ζ
))
W
d
2
W
d
2
ζ
=
2
(
u(
ζ
)F(V(
ζ
))
)
2
W
dW
d
ζ
,
dW
d
ζ
+
2
(
u(
ζ
)F(V(
ζ
))
)
2
W
d
2
W
d
2
ζ
, (2.94)
Set
ζ
=
ζ
+
and denote W(
ζ
+
)=W
+
. Notice that since
+
· [G] = 0 and
ˆ
v
+
has a
critical point at V
+
, by Proposition 2.3.6 the equality dv
s
/d
ζ
= 0 is satisfied at V
+
and dW/d
ζ
= r := r
+
.So(2.94) reduces to:
[G]
d
2
v
s
d
2
ζ
+
λ
2
G
2
W
(
r,r
)
+
G
W
d
2
W
d
2
ζ
=
2
(
uF
)
2
W
(
r,r
)
+
2
(
uF
)
2
W
d
2
W
d
2
ζ
, (2.95)
We recall that Eq.
(2.30) can be written as:
λ
G
W
r
=
(
uF
)
W
r. (2.96)
Differentiating
(2.96) with respect to
ζ
along a rarefaction curve starting at (V
+
,u
+
),
we obtain:
d
λ
d
ζ
G
W
r
+
λ
2
G
2
W
(
r,r
)
+
λ
G
W
dr
d
ζ
=
2
(
uF
)
2
W
(
r,r
)
+
(
uF
)
W
dr
d
ζ
, (2.97)
where d
λ
/d
ζ
=
λ
·r and we set
ζ
=
ζ
+
in (2.97). Manipulations of Eqs. (2.95) and
(2.97) yield:
[G]
d
2
v
s
d
2
ζ
G
W
r
+
u
λ
G
W
λ
·r
(
uF
)
W

d
2
W
d
2
ζ
dr
d
ζ
= 0 (2.98)
We multiply
(2.98) by the left eigenvector . Notice that the last term in (2.98) van-
ishes, i.e.:
·
u
λ
G
W
(
uF
)
W

d
2
W
d
2
ζ
dr
d
ζ
= 0, (2.99)
39
because from Lemma 2.2.1, is eigenvector associated to
ˆ
λ
+
. The system (2.98) re-
duces to:
·[G]
d
2
v
s
d
2
ζ
·
G
W
r
λ
·r = 0, (2.100)
after some manipulations, we see that the inequalities
(2.92) and (2.93) are satisfied.
Notice that the inequalities
(2.92) and (2.93) do not depend on the speed u because:
i)- The matrix
W
G
+
has the form (2.28) and by Lemma 2.2.1, the right eigenvec-
tors r have the form
(2.36.a), so the expression
W
G
+
r is independent of u.
ii)- From Lemma 2.2.1, we know that
does not depend on the speed u,so ·
W
G
+
r
does not depend on u.
Since G is solely a function of V, the claim follows.
Lemma 2.3.2. Assume that the shock and rarefaction curves can be parametrized by
ζ
in a neighborhood B
j
δ
(V
+
,u
+
), with
ζ
=
ζ
+
at (V
+
,u
+
). Assume also that G is smooth,
V
is sufficiently close to V
+
and
ˆ
v
+
(V
;V
+
)=
ˆ
λ
+,p
(V
+
) for some family p for
λ
,
and r, then we approximate
+
·[G] by:
+
·[G] A(
ζ
;
ζ
+
)=
+
W
G
+
r
+
(
ζ
+
ζ
), (2.101)
where V
= V(
ζ
), V
+
= V(
ζ
+
) and |
ζ
+
ζ
| <
.
Proof: Since G is smooth, we can approximate
[G] by:
[G]=
W
G
+
dW
d
ζ
(
ζ
+
ζ
)+O
(
ζ
+
ζ
)
2
.
Disregarding terms of order 2, since
ˆ
v
+
(V
;V
+
)=
ˆ
λ
+
(V
+
), by Proposition 2.3.6 it fol-
lows that dW
(
ζ
+
)/d
ζ
= r(V
+
,u
+
), so Eq. (2.101) is satisfied and the Lemma follows.
Notice that it is not necessary to know u
and u
+
, because (2.101) does not depend
on the speed u.
Proposition 2.3.8. Assume the hypotheses of Proposition 2.3.7. Then the following
relationships are valid:
i) If
A(
ξ
,
ξ
+
) > 0, then
ˆ
v
+
has a minimum.
ii) If
A(
ξ
,
ξ
+
) < 0, then
ˆ
v
+
has a maximum.
Proof: First, let V
+
and V
satisfy hypotheses of Lemma 2.3.2. From Eqs. (2.100)
and (2.101), we know that d
2
v
s
/d
2
ζ
, satisfies:
d
2
v
s
d
2
ζ
+
B
+
r
+
+
B
+
r
+
(
ζ
+
ζ
)
λ
·r
+
=
λ
·r
+
ζ
+
ζ
, (2.102)
Observe that if
λ
·r
+
> 0, then by the geometrical compatibility it is necessary that
ζ
+
>
ζ
. On the other hand, if
λ
· r
+
< 0, then
ζ
+
<
ζ
, by the same argument.
Using Eq.
(2.102), we obtain that d
2
v
s
/d
2
ζ
> 0, i.e., v
s
(and
ˆ
v
+
) is minimum at V
+
.
40
Now let V
and V
+
be such that A(
ξ
,
ξ
+
) > 0. Since
+
B
+
r
+
λ
· r
+
depends
solely on V
+
, it is necessary to analyze only
+
·[G]. Since A(
ξ
,
ξ
+
) > 0,
+
·[G] has
the same sign as
+
B
+
r
+
(
ζ
+
ζ
), then d
2
v
s
/d
2
s > 0 and v
s
(and so
ˆ
v
+
) is minimum
at V
+
.
Now let V
and V
+
be such that A(
ξ
,
ξ
+
) < 0, then
+
·[G] and
+
B
+
r
+
(
ζ
+
ζ
)
have opposite signs, so d
2
v
s
/d
2
s < 0 and v
s
(and
ˆ
v
+
) is maximum at V
+
.
Secondary bifurcation manifold
The RH locus is obtained implicitly by
H
l
(V
,V
+
)=0 for l ∈C
RH
, where H
l
: R
2n
−→
R
2n
, and V
and V
+
are the primary variables. Notice that RH is 1-dimensional lo-
cus, defined by expressions
(2.59). However, sometimes these implicit expressions fail
to define a curve for V
+
in the different physical situations, typically the curve self-
intersects. We call these points V
+
secondary bifurcation and, from the implicit
function theorem, they are the
(+) states for which there exists a () state such that
the following equalities are satisfied:
H
l
= 0 for l ∈{
1,
2,···,
−−→
n 1} (2.103)
and
J
H
=
H
l
V
+
does not have maximal rank for all l ∈{
1,
2,···,
−−→
n 1}, (2.104)
where
H
l
is defined implicitly by (2.57) and J
H
is the Jacobian matrix of the vectors
H
l
for l ∈{
1,
2,···,
−−→
n 1} corresponds to a triplet of C
RH
. The Jacobian is an
n
×
(
n 1
)
matrix in the following form:
H
1
V
1
H
1
V
2
···
H
1
V
n
H
2
V
1
H
2
V
2
···
H
2
V
n
.
.
.
.
.
.
.
.
. s
H
n1
V
1
H
n1
V
2
···
H
n1
V
n
(2.105)
The Jacobian does not have maximal rank if any two minors of size
(
n 1
)
×
(
n 1
)
have determinant zero. Following the same ideas to obtain RH locus, we can prove
that:
Corollary 2.3.3. The secondary bifurcation is an
(
n 1
)
dimensional structure in the
space of n variables V
+
, i.e., it is generically codimension1 in the space V
+
.
41
2.4 Regular RH locus
Definition 2.4.1. Given a state V
∈P
j
for j ∈PS, we say that the RH locus for V
is regular if for all k ∈PS(including k = j) and for each V
+
∈P
k
with V
+
= V
the shock speed v
s
and the speed u are finite and there are at least two distinct indices
p,q
∈Cdepending on k and j such that:
1.
F
q
[G
p
] F
p
[G
q
] = 0; (2.106)
2.
[G
q
] = 0 and the following limit exist:
lim
V
+
−→V
[G
p
]
[G
q
]
. (2.107)
3. The following inequality is satisfied:
F
p
= F
q
lim
V
+
−→V
[G
p
]
[G
q
]
. (2.108)
Remark 2.4.1. When V
is a frontier point, then the limits in (2.107) and (2.108) may
not be well defined. However, since we are interested in obtaining the solution in the
physical situation
P
j
we regard the limits as lateral limits.
2.4.1 Small shocks
A shock between two states W
and W
+
is said to be small if there exists a sufficiently
small
> 0 such that |W
+
W
| <
. For this type of shock we can prove important
results that can be extended to more general shocks.
Speed for regular small shocks
Lemma 2.4.1. Assume that F and G are
C
2
for V ∈P
j
. In a regular system, when the
primary variable V
+
converges to V
the speed u
+
converges to u
.
Proof:
Let V
be fixed. From eq. (2.63.b), we define for all V
+
∈P
j
the expression:
u
+
= u
F
q
[G
p
] F
p
[G
q
]
F
+
q
[G
p
] F
+
p
[G
q
]
. (2.109)
We can rewrite
(2.109) as:
(u
+
u
)
F
q
[G
p
] F
p
[G
q
]
= u
+
[F
q
][G
p
] [F
p
][G
q
]
. (2.110)
42
Assume that [G
q
] = 0. Dividing (2.110) by [G
q
], we obtain:
(u
+
u
)
F
q
[G
p
]
[G
q
]
F
p
= u
+
[F
q
]
[
G
p
]
[G
q
]
[
F
p
]
. (2.111)
Let V
+
∈RH(V
) tend to V
and remember that the limit (2.107) exists and (2.108)
is satisfied. From Eqs. (2.106)-(2.108), u
+
is bounded. Notice that [F
q
] and [F
p
] tend
to zero when V
+
tends to V
,sou = u
+
= u
.
Extension of Lax theory
Often shocks between regions are not weak and the cumulative and flux terms are not
smooth; the Lax theory cannot be extended to this type of shocks. However, within
a physical situation and under certain hypotheses we can obtain a theory for the
Riemann solution similar to Lax’s. We consider a physical situation j
∈PS.
We say that the system
(2.2) represents a strictly hyperbolic physical situation if
there are n distinct eigenvalues
λ
p
= u
ϑ
p
(V) that satisfy:
ϑ
1
(V) <
ϑ
2
(V) < ···<
ϑ
n
(V). (2.112)
We know that if
(2.112) is satisfied, there exist n right r
p
(and n left
p
) eigenvectors
that are LI.
From
(2.36) of Lemma 2.2.1, the right and left eigenvectors of the system (2.2) have
the form r
p
=(g
p
1
(V), ··· , g
p
n
(V), ug
p
n
+1
(V)) and
p
=(
p
1
(V), ··· ,
p
n
(V),
p
n
+1
(V)). No-
tice that neither the right nor the left vectors form bases for the space
(V,u).
Nevertheless, the left and right eigenvectors satisfy the following lemma:
Lemma 2.4.2. Consider
λ
i
=
λ
j
, and the associate right and left eigenvectors r
i
and
j
, then:
j
W
Gr
i
= 0. (2.113)
We can prove now that if the system
(2.2) is satisfied in the regular case there
exists a local basis of shock curves in the space of primary variables in each physical
situation.
Proposition 2.4.1. Let W
=(V
,u
) be any point in phase space with V
in the
interior of
P
j
. Then there are n C
2
one-parameter curves; of states V
k
= V
k
(
θ
) for
k
= 1,2,···,n and
θ
sufficiently small, where V
k
(0)=V
. The points V
k
(
θ
) satisfy the
RH condition
(2.52). Moreover the following relationships are satisfied:
V
k
(
θ
)=V
+
θ
ˆ
r
k
(V
)+
θ
2
2
V
ˆ
r
k
(V
)
ˆ
r
k
(V
)+O(
θ
3
) (2.114)
and
v
s
k
=
λ
k
(W
)+
θ
2
2
W
λ
l
(W
) ·r
k
(W
)+O(
θ
2
). (2.115)
We can also obtain the speed u
k
= u
k
(V(
θ
)) from V
k
(
θ
) and Eq. (2.64.b),fork =
1,2,··· ,n.
43
Proof: We follow the same idea of the original theorem of Lax [43]. We can write:
v
s
(G(V) G(V
)) =
1
0
d
d
σ
(G(V
+
σ
(V V
)))d
σ
=
1
0
V
G(V
+
σ
(V V
))(V V
)d
σ
(2.116)
Similarly, from Eq.
(2.64.a), we can rewrite uF(V) u
F(V
). Using u = u
χ
, with
χ
:=
χ
(V
,V) we define
ˆ
W
ˆ
W
=
(
V V
,
χ
1
)
, and F as follows:
uF
(V) u
F(V
)=u
1
0
d
d
σ
(1 +
σ
(
χ
1)) F(V
+
σ
(V V
))
d
σ
= F(
ˆ
W
ˆ
W
), (2.117)
where
F := F(W
,W
+
) is an (n + 1) ×(n + 1) matrix given by:
F := u
1
0
(1 +
σ
(
χ
1))
V
F(V
+
σ
(V V
)), F(V
+
σ
(V V
))
d
σ
.
and
ˆ
W
ˆ
W
=
(
V V
,
χ
1
)
.
From Eq.
(2.116) we can see that v
s
(G(V) G(V
)) does not depend on u. How-
ever, it is necessary rewrite it in the variable
ˆ
W, so we define the
(n + 1) × (n + 1 )
matrix G(V) as:
G(V) :=
1
0
V
G(V
+
σ
(V V
),0)
d
σ
.
Thus, we can rewrite v
s
(G(V) G(V
)) as:
v
s
(G(V) G(V
)) = v
s
G(V)(
ˆ
W
W
). (2.118)
Using Eqs.
(2.117) and (2.118), the Rankine-Hugoniot condition (2.50) becomes:
(
F(V) v
s
G(V)
)
(V V
,
χ
1)=0. (2.119)
Using
(2.119), we know that a state V ∈Nbelongs to the RHL locus of V
, if, only
if, there exists a k
∈{1,2,··· ,n} such that:
v
s
k
(V
,u
;V)=
µ
k
(u
,V). (2.120)
and that
(V V
,
χ
1) is a right eigenvector of F(V)
µ
G(V) associated to
µ
k
(u
,V).
Notice that when V converges to V
by the Lemma 2.4.1 the speed u converges to
u
, so the matrices F(V) −→ A and G(V) −→ B, with B and A given by Eq. (2.28) and
(2.29). Since A
λ
B has n distinct eigenvalues that satisfy (2.112), and the functions
F(V) and G( V) are continuous, there exists a neighborhood of N of V
in the topology
of
P
j
and n distinct functions
µ
k
(V,V
,u
) that are the eigenvalues of F(V)
µ
G(V),
that satisfy
µ
k
(V
,V
,u
)=
λ
(V
,u
).
44
We denote the left eigenvectors of the system F( V)
µ
G(V) by (L
k
(V))
T
. Since
(L
k
(V))
T
is not a basis in (V,u) space, for (V V
,
χ
1) to be a right eigenvector,
the condition:
L
j
(V)G(V V
,
χ
1)=0 j = k (2.121)
is only necessary. Notice that the first n coordinates of the eigenvectors are uniquely
defined by Eq.
(2.121). Moreover, if V ∈Nbelongs to the RHL of V
, the speed u is
given by Eq.
(2.64.a), thus Eq. (2.121) defines uniquely the eigenvectors of F(V)
µ
G(V).
Since
G has a zero column, Eq. (2.121) does not depend on
χ
1, and Eq. (2.121) is
a system with
(n 1) equations in n unknowns V. This system which can be written
as:
Φ
k
(V)=M
k
(V)(V V
)=0, (2.122)
where
M
k
(V)=
L
1
(V)
T
.
.
.
L
k1
(V)
T
L
k+1
(V)
T
.
.
.
(
L
n
(V)
)
T
ˆ
G, (2.123)
with
ˆ
G =
1
0
V
G(V
+
σ
(V V
))d
σ
.
Notice that at V
we obtain:
Φ
k
(V
)=0 and
V
Φ
k
(V
)=M
k
(V
). (2.124)
Since L
(V
)=(V
) and
ˆ
G have rank n, the matrix M
k
(V) also rank n 1, yielding
that the matrix
V
Φ
k
(V
) has rank n 1. By the implicit function theorem, there
exists a one-parameter family of functions V
k
(
), for 1 k n and |
|≤
1
small
enough, with
V
k
(0)=V
, (2.125)
and the n different waves with shock speed satisfying:
lim
V−→V
v
s
(V
,u
;V)=
λ
k
(V
,u
) for 1 k n. (2.126)
Letting
tend to zero in Eqs. (2.121), we obtain that dV(0)/d
is parallel to
ˆ
r
k
(V
). After a change in parametrization, we obtain
dV
(0) /d
=
ˆ
r
k
(V
). (2.127)
After obtaining V, Eq.
(2.87) determines u. Differentiating (2.89), using that v
s
satisfies Eq. (2.87), from (2.127) it follows that:
dV
d
,
du
d
(0)=r
k
(V
) for k = 1,2,··· ,n. (2.128)
45
We drop the index k in the expressions. Now using Eqs. (2.94) and (2.97) we can prove
the expansion
(2.115). To do so, first we subtract (2.94) from (2.97) and set
= 0 we
obtain:
2
dv
s
d
G
W
r
G
W
r
λ
·r +
λ
G
W
(
uF
)
W

d
2
W
d
2
dr
d
= 0, (2.129)
with
λ
=
λ
(V
,u
) and r = r
k
(V
,u
) for some k = 1,2,··· ,n. Multiplying Eq.
(2.129) by and using (2.99), we obtain:
2
dv
s
d
G
W
r
=
G
W
r
λ
·r, (2.130)
Since
W
Gr = 0, we get:
dv
s
d
=
1
2
λ
·r. (2.131)
Substituting
(2.131) in Eq. (2.129) and noticing that dr/d
=
W
r ·r, we obtain:
λ
G
W
(
uF
)
W

d
2
W
d
2
dr
d
. (2.132)
Therefore, there exists a real number
α
such that
d
2
V
d
2
,
d
2
u
d
2
=
d
2
W
d
2
=
d
ˆ
r
d
+
α
ˆ
r,u
dr
n+1
d
+
α
ur
n+1
. (2.133)
Changing the parametrization by setting:
=
θ
1
2
αθ
2
. (2.134)
Then using Taylor expansion, Eqs.
(2.126) and (2.131) yield Eq. (2.114):
v
s
=
λ
+
2
λ
·r + O(
2
),
=
λ
+
θ
2
λ
·r + O(
2
).
Eq.
(2.115) is obtained by using Eq. (2.133):
V
(
)=V
+
ˆ
r
+
2
2
(
V
ˆ
r
·
ˆ
r
+
α
ˆ
r
)
+
O(
3
)
=
V
+
θ
ˆ
r
+
θ
2
2
V
ˆ
r
·
ˆ
r
+ O(
3
).
Under certain hypotheses, we can prove the Theorem of Lax to construct the local
Riemann solution within each physical situation.
46
2.4.2 Other degeneracies of RH locus
In addition to the linear degeneracy 2.3.2, we study another degeneracy that seldom
occurs. To do so, we utilize the notation:
D
pq
(V
,V
+
) := F
+
p
[G
q
] F
+
q
[G
p
].
Proposition 2.3.1 is valid if non zero denominators can be found for Eq.
(2.66). So it is
necessary to study the behavior of the solution when
D
pq
(V
,V
+
)=0 for all p , q ∈C.
For a fixed pair p, q
∈C, we can follow the same steps of Proposition 4.4.3 and prove:
Proposition 2.4.2. If
D
pq
(V
,V
+
)=0 for all p,q ∈C, it follows that:
[G
k
]=
ρ
1
F
+
k
and F
k
=
ρ
2
F
+
k
, k ∈C.
where
ρ
1
and
ρ
2
are numbers depending on [G], F
and F
+
. Moreover the shock speed
v
s
satisfies:
v
s
=
χ
(V
;V
+
)
ρ
2
ρ
1
u
,
where
χ
(V
;V
+
) is given by Eq. (2.64.a).
2.4.3 The sign of u
+
on regular Rankine-Hugoniot loci
In the previous Propositions, it is necessary that u is never zero. Moreover, it is
useful that u and the injection speed u
have the same sign. From Eq. (2.43), these
properties are clearly true on the integral curves (and so on rarefaction waves), but
they are not obvious on shock waves. We now prove a sign invariance for the connected
branches of the RHL containing
(V
,u
), that we denote by CRH L.
To do so, for a fixed V
in some P
j
, we define for each V ∈P
k
(including k = j ) and
for all index pairs p,q
∈C, the following continuous functions Ξ
pq
:= Ξ
pq
(V,V
) from
the denominator and numerator of the expression for u given in
(2.64.b):
Ξ
pq
=
F
p
(G
q
G
q
) F
q
(G
p
G
p
)

F
p
(G
q
G
q
) F
q
(G
p
G
p
)
, (2.135)
where the cumulative and flux terms depend on j and k and G
= G(V), G
= G(V
),
F
= F(V) and F
= F(V
). The shock curves can occur between different physical
situations, therefore it is useful to represent the states on RHL as belonging to
j
.
Using
(2.18), we can write Ξ
pq
:= Ξ
pq
(V
,V
+
).
Remark 2.4.2. Notice that for
V
and V
+
∈RHthere exists at least a pair p, q ∈C
such that
ˆ
X
pq
(V
,V
+
) = 0 in the regular case. By the hypothesis of regular case,
the speed u is finite, then using the Eq.
(2.64.b) we can see that the denominator of
expression
(2.64.b) is zero, if only if, its numerator is zero. We conclude that if Ξ
pq
is
zero for all p, q, then the numerator of
(2.64.b) is zero for any p, q, that is D
pq
= 0 for
all p, q. It is impossible because it violates the hypothesis 1 of Definition 2.4.1.
47
It follows from Prop. 2.1.1 that:
Lemma 2.4.3. Ξ
pq
are continuous functions for all p, q ∈C.
Remark 2.4.3. Notice that for u
+
given by (2.63), the inequality u
+
u
< 0 is satisfied
if, only if, Ξ
pq
(V
,V
+
) < 0.
Remark 2.4.4. For two fixed states
V
and V
+
∈RH(V
) for all pairs p, q ∈Ceither
Ξ
pq
(V
,V
+
) 0 or Ξ
pq
(V
,V
+
) 0. This fact occurs because for points in RH locus
the speed is uniquely defined.
Lemma 2.4.4. For
V
j
, the expression Ξ
pq
given by Eq. (2.135) satisfies:
1. Ξ
pq
= 0 for V
+
= V
and for all p, q ∈C;
2. There exists an
> 0 and and indices p,q ∈Csuch that for V∈CRHL
B
j
(V
),
the inequality Ξ
pq
(V
,V
+
) > 0 is satisfied.
Proof: The assertion
(1) is immediate. Because u
+
= u
when V
+
converges
to
V
, there exists an
> 0 such that for each V∈CRHL
B
(V
), it follows that
Ξ
pq
(V
,V) 0. Moreover, since we assume that the system is regular in the sense of
Definition 2.4.1, for each V
B
(V
)
CRH L\V
there exists at least a pair p,q ∈C
such that Ξ
pq
(V
,V) > 0, and (2) is valid.
Definition 2.4.2. For fixed V
j
, we utilize Ξ
pq
given in (2.135) to define the two
disjoint sets:
J
,0
:=
V
k
CRH L
k
, for all k = PS, such that for all p, q ∈C
Ξ
pq
(V
,V
l
) 0, but there is at least a pair l,m ∈C, such that Ξ
lm
(V
,V
k
) < 0
(2.136)
and
J
+,0
:=
V
k
CRH L
k
, for all k ∈PS, such that for all p,q ∈C
Ξ
pq
(V
,V
k
) 0, but there is at least a pair l,m ∈C, such that Ξ
lm
(V
,V
k
) > 0
(2.137)
These two sets form a disjoint decomposition of CRHL in as summarized in the
Lemma below:
Lemma 2.4.5.
J
,0
J
+,0
= and J
,0
J
+,0
= CRH L(V
)\V
.
Proof: Let a
V∈J
,0
J
+,0
.IfV∈J
,0
then Ξ
pq
(V
,V) 0 for all p, q ∈Cand
there exists at least a index pair l, m such that Ξ
lm
(V
,V) < 0; similarly, if V∈J
+,0
,
then there exists a index pair i, j
∈Csuch that Ξ
ij
(V
,V) > 0. From Remark 2.4.4,
we know that it does not occurs, so
J
,0
J
+,0
= .
Consider now
V∈CRH L(V
). It is immediate that V belongs to the J
,0
J
+,0
.
Notice that
V
there is no in J
,0
J
+,0
because of Lemma 2.4.4.
From the continuity of Ξ
pq
(V
,V) we obtain the following lemma.
48
Lemma 2.4.6. The sets J
,0
and J
+,0
are connected and open in CRHL.
Since CRHL is connected and from Lemma 2.4.4 we know that there is a pair
p,q
∈Csuch that Ξ
pq
(V
,V) for V belongs to the CRHL(V
) in a neighborhood, the
Lemmas 2.4.5 and 2.4.6 we yield:
Proposition 2.4.3. Let any
V
j
. The speed u has the same sign of the speed u
for the state V
on the CRHL.
We show some applications of the theory developed in Appendix C.
49
CHAPTER 3
Mathematical modelling of thermal oil recovery with
distillation
We present a physical model for steam and gaseous volatile oil injection into a hori-
zontal, linear porous medium filled with oil and water. The oil in the porous medium
consists of dead oil and dissolved volatile oil. Dead oil is an oil with negligible va-
por pressure while volatile oil is an oil with substantial vapor pressure. The model
is based on mass balance and energy conservation equations as well as on Darcy’s
law. We present the main physical definitions and summarize the equations of the
model. This system of equations is a representative example of flow in porous media
with mass interchange between phases (in the absence of gravitational effects). Such
flows are not represented by conservation laws or hyperbolic equations both because
they have source terms and because they are not evolutionary in all the variables.
The model contains a system of 6 equations: 5 equations for mass balance and one
equation for energy conservation, which is generically represented by:
t
G(V)+
x
u
F(V)=Q(V) , (3.1)
where
V =(s
w
,s
g
, T,v
ov
, y
gv
) R
5
represents the 5 dependent variables to be de-
termined for all x and t
0. The cumulative and flux terms are G = {G
1
,G
2
,···,G
6
}
and F = {F
1
,F
2
,···,F
6
}. Here s
w
and s
g
are the water and gas saturations, T is
the temperature, v
ov
and y
gv
are the volatile oil volume fraction in the oleic gaseous
phases, respectively. The real-valued dependent variable u is the Darcy speed. In
(3.1), uF
i
is the flux for the conserved quantity G
i
and G
i
/t is the corresponding
accumulation term, in each of the 6 equations. The source term Q
= {Q
1
, Q
2
,···, Q
6
}
represents the water and volatile oil mass transfer between the phases; the compo-
nent Q
6
vanishes because it represents the conservation of energy total. The functions
G
i
and F
i
are continuous in the whole domain , but later we will see that they are
only piecewise smooth. In the
(x, t) plane the terms Q often generate fast variations
50
in the variables V(x, t), so that Q(V(x, t)) may be regarded as distribution. Eq. (3.1)
has another important feature: the variable u does not appear in the accumulation
term, but only in the flux term.
Thermodynamics plays a very important role in the physical phenomena modelled
by Eq.
(3.1). The thermodynamical behavior of interest can be divided into systems in
equilibrium or in quasi-equilibrium. For such types of behavior we can use the Gibb’s
phase rule, which determines the number of thermodynamical degrees of freedom:
f
= c p + 2, (3.2)
where f is the number of thermodynamical degrees of freedom, c is the number of
components and p is the number of phases. Such flows encompass particular physi-
cal situations, where one or several phases or chemical components are missing. We
study the possible physical component and phase mixture situations in thermody-
namical equilibrium or quasi-equilibrium that appear in the Riemann solution. For
equilibrium situations, Q
= 0. In the situations in quasi-equilibrium Q is not zero
but very small, so that the thermodynamical degrees of freedom still satisfy Eq.
(3.2)
in all physical situations. Finally, Q has fast variation in the regions between the
physical situations. In the model considered in this work, there is a map E that can
be applied to the balance system
(3.1). This map reduces (3.1) to simpler systems of
conservation laws, that is, E eliminates the terms Q; F and G are obtained from
F and
G through E and reduces to equation of form (2.2) After this elimination, further sim-
plifications may occur in each physical situation, so that each situation is described
by a set of variables V that is a subset of the set of variables
V. Each of the systems
of type
(2.2) has fewer equations than the complete system (3.1). It is useful to define
W
=(V, u). Although Eq. (2.2) is a conservation law, it is not a hyperbolic system, be-
cause the variable u appears solely in the flux term. Indeed, Eq.
(2.2) has an infinite
speed mode associated to u. Under certain hypotheses we show that if u
> 0 the Rie-
mann solution in the variables
V does not depend on u. Thus the unknown variables
in the system of equation for each physical situation are called primary variables,to
be determined from the corresponding version of (2.2). Because the variable u can be
obtained from the primary variables, we call it a secondary variable. The variables
that are constant or that are obtained in a simple way from the primary variables are
called trivial variables.
3.1 Physical and Mathematical Models
We consider the injection of steam and volatile oil into a linear horizontal porous rock
cylinder with constant porosity and absolute permeability, containing water and oil
initially. The oil consists of dead oil with or without volatile oil. There are three chem-
ical components: water, volatile oil and dead oil,
(w,v, d) as well as three immiscible
phases: aqueous (liquid water), oleic (liquid oil) and gaseous,
(w,o, g). We use the
following convention in subscripts for physical properties: the first subscript
(w,o, g)
refers to the phase and the second subscript (w,v,d) refers to the component.
51
We use simple mixing rules. We disregard any heat of mixing both in the gaseous
phase between steam and volatile oil and in the oleic phase between volatile oil and
dead oil. Moreover we disregard any volume contraction effects resulting from mixing.
The concentration
kg
/m
3
of (dead) volatile oil in the oleic phase is denoted as
(
ρ
od
)
ρ
ov
. In the same way we define the concentration of the volatile oil (water vapor) in
the gaseous phase as
ρ
gv
ρ
gw
. For ideal fluids we obtain
ρ
ov
ρ
V
+
ρ
od
ρ
D
= 1,
ρ
gw
ρ
gW
+
ρ
gv
ρ
gV
= 1, (3.3)
where the capital subscripts indicates that the component is in a pure phase. The
pure liquid densities of volatile and dead oil in the oleic phase, denoted by
ρ
V
,
ρ
D
[kg/m
3
], are considered to be independent of temperature; the pure vapor densities of
steam and volatile oil in the gaseous phase, denoted by
ρ
gW
and
ρ
gV
, are considered
to obey the ideal gas law, i.e.,
ρ
gW
=
M
W
P
RT
,
ρ
gV
=
M
V
P
RT
, (3.4)
where M
W
, M
V
denote the molar weights of water and volatile oil respectively. T is
the variable temperature and the gas constant is R
= 8.31[J/mol /K]. The pressure P
is not a variable in this problem, but the fixed prevailing pressure value.
We define x
ov
as the mole fraction of volatile oil in the oleic phase and y
gv
the
mole fraction of volatile oil in the vapor phase. In some physical situations, in addi-
tion to temperature and pressure, it is necessary to specify another thermodynamical
variable: either x
ov
or y
gv
, which are explained in the following text.
Using Clausius-Clapeyron law, we can write the the pure water vapor pressure P
w
as:
P
w
(
T
)
=
P
o
exp
M
W
Λ
W
T
w
b
R
1
T
1
T
w
b
, (3.5)
where Λ
W
T
w
b
[
J/kg
]
is its evaporation heat at its normal boiling temperature T
w
b
[
K
]
,
at P
o
, the atmospheric pressure. Eq. (3.5) is valid for any temperature in situations
where there exists liquid water.
We also use Clausius-Clapeyron law for the vapor pressure of pure volatile oil:
P
V
(
T
)
=
P
o
exp
M
V
Λ
V
T
v
b
R
1
T
1
T
v
b
, (3.6)
where Λ
V
T
v
b
is its evaporation heat at its normal boiling temperature T
v
b
and atmo-
spheric pressure. Eq.
(3.6) is valid for any temperature below T
v
b
. In addition, we use
Raoult’s law, which states that the vapor pressure of volatile oil is equal to its pure
vapor pressure given by Eq. (3.6
) times the mole fraction x
ov
of volatile oil dissolved
in the oil phase. Therefore we obtain
P
v
(
T
)
=
x
ov
P
o
exp
M
V
Λ
V
T
v
b
R
1
T
1
T
v
b
. (3.7)
52
We assume that the prevailing pressure P is the sum of the two vapor pressures: P
w
and P
v
.
We can use Eq. (3.3.a) and define the volatile oil volume fraction and the dead oil
volume fraction in the oleic phase respectively, as:
v
ov
=
ρ
ov
/
ρ
V
,1v
ov
=
ρ
od
/
ρ
D
,
(3.8)
Similarly, we can use Eq. (3.3.b) and define the volatile oil volume fraction and the
steam water in the gaseous phase respectively, as:
y
gv
=
ρ
gv
/
ρ
gV
,1 y
gv
=
ρ
gw
/
ρ
gW
,
(3.9)
We can derive an equation that relates the oleic phase density to the mole frac-
tion x
ov
. From the definition of the mole fraction (moles volatile oil / total moles) we
immediately observe that
x
ov
=
ρ
ov
/M
V
ρ
ov
/M
V
+
ρ
od
/M
D
. (3.10)
The concentration of steam and volatile oil in the gaseous phase is given by:
ρ
gw
=
M
W
P
w
RT
,
ρ
gv
=
M
V
P
v
RT
. (3.11)
In the physical situations in which there exists volatile oil in the oleic phase, we can
use Eqs.
(3.7), (3.9. a) and (3.11.b), we derive the following expression for the mole
fraction of volatile oil in the vapor phase y
gv
as a function of x
ov
and T:
y
gv
(
T, x
ov
)
=
P
v
(
T
)
P
=
P
o
P
x
ov
exp
M
V
Λ
V
T
v
b
R
1
T
1
T
v
b
. (3.12)
We assume that the rock is filled with a mixture of water, oil and gas, i.e.:
s
w
+ s
g
+ s
o
= 1, (3.13)
where s
w
, s
g
and s
o
are the oil, gas and water saturations, i.e., the pore volume fraction
is filled with the liquid water, gaseous and oleic phases respectively. We usually write
s
o
in terms of s
g
and s
w
.
We assume that the fluids are incompressible and that the pressure changes are
so small that they do not affect the physical properties of the fluids. Darcy’s law for
multiphase flow relates pressure gradient with its Darcy speed:
u
w
=
kk
rw
µ
w
p
x
, u
g
=
kk
rg
µ
g
p
x
and u
o
=
kk
ro
µ
o
p
x
. (3.14)
The relative permeability functions k
rw
, k
rg
and k
ro
are considered to be functions of
the saturations (see Appendix A);
µ
w
,
µ
g
and
µ
o
are the viscosities of the aqueous,
gaseous and oleic phases. The pressure in the aqueous, gaseous and oleic phases is
taken to be identical, p, because the capillary pressure between the different phase
53
is neglected. The fractional flow functions for water phase, gaseous phase and oleic
phase are defined by:
f
w
=
k
rw
/
µ
w
d
, f
g
=
k
rg
/
µ
g
d
and f
o
=
k
ro
/
µ
o
d
, (3.15)
where d
= k
rw
/
µ
w
+ k
rg
/
µ
g
+ k
ro
/
µ
o
.
Using Darcy’s law (3.14) and Eq. (3.15):
u
w
= uf
w
, u
g
= uf
g
and u
o
= uf
o
, where u = u
w
+ u
g
+ u
o
(3.16)
is the total or Darcy speed.
3.1.1 Balance Equations.
We can write the mass balance equations for liquid water, steam in the gaseous phase,
volatile oil in the gaseous phase, volatile oil in the oleic phase and dead oil conserva-
tion respectively as:
ϕ
t
ρ
W
s
w
+
x
ρ
W
u
w
= q
g−→ a,w
, (3.17)
ϕ
t
ρ
gw
s
g
+
x
ρ
gw
u
g
= q
g−→ a,w
, (3.18)
ϕ
t
ρ
gv
s
g
+
x
ρ
gv
uf
g
= q
g−→o,v
, (3.19)
ϕ
t
ρ
ov
s
o
+
x
ρ
ov
u
o
= q
g−→o,v
, (3.20)
ϕ
t
ρ
od
s
o
+
x
ρ
od
u
o
= 0, (3.21)
where q
g−→ a,w
and q
g−→o,v
are the water and the volatile oil condensation rates. The
steam condenses into liquid water and the volatile oil dissolves into the dead oil.
The conservation of energy in terms of enthalpies per unit mass is given as
t
H
r
+
ϕ
s
w
ρ
W
h
W
+
ϕ
s
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
ϕ
s
g
ρ
gw
h
gW
+
ρ
gv
h
gV

+
x
u
w
ρ
W
h
W
+ u
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
u
g
ρ
gw
h
gW
+
ρ
gv
h
gV

= 0. (3.22)
Using u
w
, u
g
and u
o
given by Eq. (3.16) in Eqs. (3.17)-(3.22), we obtain:
54
ϕ
t
ρ
W
s
w
+
x
u
ρ
W
f
w
= q
g−→ a,w
, (3.23)
ϕ
t
ρ
gw
s
g
+
x
u
ρ
gw
f
g
= q
g−→ a,w
, (3.24)
ϕ
t
ρ
gv
s
g
+
x
u
ρ
gv
f
g
= q
g−→o,v
, (3.25)
ϕ
t
ρ
ov
s
o
+
x
u
ρ
ov
f
o
= q
g−→o,v
, (3.26)
ϕ
t
ρ
od
s
o
+
x
u
ρ
od
f
o
= 0, (3.27)
t
H
r
+
ϕ
s
w
ρ
W
h
W
+
ϕ
s
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
ϕ
s
g
ρ
gw
h
gW
+
ρ
gv
h
gV

+
x
u
f
w
ρ
W
h
W
+ f
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
f
g
ρ
gw
h
gW
+
ρ
gv
h
gV

= 0. (3.28)
3.1.2 Reduced system of equations
The system of balance laws ( 3.23)-(3.28) is written in the conservative form by adding
(3.23) to (3.24) and (3.25) to (3.26):
ϕ
t
ρ
W
s
w
+
ρ
gw
s
g
+
x
u
ρ
W
f
w
+ f
g
ρ
gw
= 0, (3.29)
ϕ
t
ρ
gv
s
g
+
ρ
ov
s
o
+
x
u
ρ
gv
f
g
+
ρ
ov
f
o
= 0, (3.30)
ϕ
t
ρ
od
s
o
+
x
ρ
od
uf
o
= 0,
t
H
r
+
ϕ
s
w
ρ
W
h
W
+
ϕ
s
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
ϕ
s
g
ρ
gw
h
gW
+
ρ
gv
h
gV

+
+
x
u
f
w
ρ
W
h
W
+ f
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
f
g
ρ
gw
h
gW
+
ρ
gv
h
gV

= 0.
Eq.
(3.29) is the conservation of component water in the steam and gaseous
phases; Eq.
(3.30) is the conservation of volatile oil in the gaseous and oleic phases.
Notice that it is possible to substitute Eq.
(3.27) by the the conservation of total oil,
which is obtained by adding Eqs.
(3.25)-( 3.27):
ϕ
t
ρ
gv
s
g
+(
ρ
od
+
ρ
ov
)s
o
+
x
u
ρ
gv
f
g
+(
ρ
ov
+
ρ
od
) f
o
= 0 (3.31)
The map E is applied to the cumulative and flux vectors generically given by
G
and F in Eq. (3.1). There is one such map E from (3.23)-(3.28) to ( 3.29), (3.30)(3.27),
55
(3.28) and another to (3.29), (3.30), (3.31), (3.28). They are:
E
1
=
110000
001100
000010
000001
and E
2
=
110000
001100
001110
000001
. (3.32)
The isomorphism (basis change) taking E
1
to E
2
and its inverse taking E
2
to E
1
are:
100000
010000
001000
000100
001110
000001
,
10 0 0 00
01 0 0 00
00 1 0 00
00 0 1 00
00
1 110
00 0 0 01
.
In this work we will use E
1
.
3.2 Physical situations
This model exhibits a rich structure of physical situations depending on the injection
and initial conditions. Although certain physical situations seldom occur, for com-
pleteness we mention all of them. In the formalism proposed in Chapter
[2 ], we can
identify 12 physical situations
j
, 12 space of primary variables P
j
and the 12 space
of trivial variables
P
j
for j = I, II,···, XII.
3.2.1 Three phases.
In this case we have three phases and three components. Using Gibb’s phase rule
(3.2), we see that there are two thermodynamical degrees of freedom. Since the pres-
sure is fixed, the only degree of freedom is the temperature. Thus x
ov
and y
gv
are
uniquely defined by the temperature T, as explained now.
From Eqs. (3.5), (3.7) and P
= P
w
(
T
)
+
P
v
(
T
)
, for the mole fraction in the liquid
phase x
ov
we find an expression depending on the temperature:
x
ov
(
T
)
=
ρ
ov
M
V
ρ
ov
M
V
+
ρ
od
M
D
=
P P
o
exp
M
W
Λ
W
(
T
w
b
)
R
1
T
1
T
w
b
P
o
exp
M
V
Λ
V
(
T
v
b
)
R
1
T
1
T
v
b
. (3.33)
Since x
ov
is function of temperature, we use Eq. (3.8.a) to obtain v
ov
as a function
of temperature only:
v
ov
=
x
ov
ρ
D
M
V
x
ov
ρ
D
M
V
+(1 x
ov
)
ρ
V
M
D
, (3.34)
56
where x
ov
is obtained from expression (3.33). The system of equations reduces to
(3.29), (3.30)(3.27), (3.28) (alternatively to the system (3.29), (3.30), (3.31), (3.28) ).
The mole fraction of volatile oil in the vapor phase y
gv
is obtained using Eq. (3.12) .
The primary variables to be determined are s
w
, s
g
, T,soP
I
= {s
w
,s
g
, T}; the speed
u is not constant, it is the secondary variable; and the trivial variables are v
ov
and y
gv
,
so
P
I
= {v
ov
, y
gv
}. The physical situation is characterized by
I
= {s
w
,s
g
, T,v
ov
, y
gv
}.
Notice for this physical situation i
I
(s
w
,s
g
, T,v
ov
, y
gv
) :=(s
w
,s
g
, T); the terms R
T
G
and R
T
F
are functions with 6 components given by:
R
T
G
=
ϕρ
W
s
w
,
ϕρ
gw
s
g
,
ϕρ
gv
s
g
,
ϕρ
ov
s
o
,
ϕρ
od
s
o
,
H
r
+
ϕ
s
w
ρ
W
h
W
+
ϕ
s
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
ϕ
s
g
ρ
gw
h
gW
+
ρ
gv
h
gV

.
and
R
T
F
=
ρ
W
f
w
,
ρ
gw
f
g
,
ρ
gv
f
g
,
ρ
ov
f
o
,
ρ
od
f
o
,
f
w
ρ
W
h
W
+ f
o
(
ρ
ov
h
oV
+
ρ
od
h
oD
)
+
f
g
ρ
gw
h
gW
+
ρ
gv
h
gV

.
3.2.2 Two phases
In this section, the number of phases is two, i.e., p = 2 in Gibb’s phase rule (3.2).
Steam phase and liquid water phase.
There is one component. Using Gibb’s phase rule
(3.2), we can see that T = T
w
b
is
constant. There is no oil, so v
ov
is not defined a priori and y
gv
= 0. If it is necessary to
obtain the complete solution, we define the variables using limits in the variable v
ov
from the contiguous situations.
The primary variable to be determined is s
w
,soP
II
= {s
w
}; the speed u is constant;
the trivial variables are s
g
= 0, T = T
w
b
, y
gv
= 0 and v
ov
,soP
II
= {s
g
= 0,T =
T
w
b
, y
gv
= 0,v
ov
}. The physical situation is characterized by
II
= {sw, s
g
= 0,T =
T
w
b
,v
ov
, y
gv
= 0}.
The system of equations reduces to a single conservation law, Buckley-Leverett’s
equation:
ϕ
t
s
w
+ u
x
f
w
= 0. (3.35)
Notice for this physical situation i
I
(s
w
,s
g
, T,v
ov
, y
gv
) :=(s
w
); the terms R
T
G
and R
T
F
are functions with 6 components given by:
R
T
G
=
ϕρ
W
s
w
,
ϕρ
gw
s
g
,0,0,0,H
r
+
ϕ
s
w
ρ
W
h
W
+
ϕ
s
g
ρ
gw
h
gW
.
and
R
T
F
=
ρ
W
f
w
,
ρ
gw
f
g
,0,0,0,f
w
ρ
W
h
W
+ f
g
ρ
gw
h
gW
.
57
Gaseous and oleic phases.
There are three components, water, volatile and dead oil. Using Gibb’s phase rule
(3.2), in addition to the temperature T, we need to find another thermodynamical
variable, v
ov
. However, notice that we could utilize as a primary variable the quantity
v
od
= 1 v
ov
. Generically, there is no unique choice between primary and trivial
variables.
Inverting Eq.
(3.34), we obtain x
ov
as function of v
ov
:
x
ov
=
v
ov
ρ
v
M
D
v
ov
ρ
v
M
D
+(1 v
ov
)
ρ
D
M
V
. (3.36)
In this physical situation, y
gv
is given by Eq. (3.12) as a function of T and x
ov
. The
variable y
gv
can be obtained using Eqs. (3.36), (3.8.a) and (3.8.b); notice that x
ov
can
be written as function of v
ov
. Thus from Eq. (3.12) y
gv
depends on v
ov
and T.
The primary variables to be determined are s
g
, T, v
ov
.soP
III
= {s
g
, T,v
ov
};
the speed u is not constant; finally the trivial variables are s
w
= 0 and y
gv
,so
P
III
= {s
w
= 0, y
gv
(T,v
ov
)}. The physical situation is characterized by
III
= {s
w
=
0,s
g
, T,v
ov
, y
gv
(T,v
ov
)}.
Using Eqs.
(3.9.a) and (3.9.b), we can rewrite the system (3.29), (3.30) , (3.27),
(3.28) in the variables s
g
, T, v
ov
and u , and obtain the following system of conservation
for steam, volatile oil, dead oil and energy:
ϕ
t
ρ
gw
s
g
+
x
uf
g
ρ
gw
= 0, (3.37)
ϕ
t
ρ
gv
s
g
+
ρ
V
v
ov
s
o
+
x
u
ρ
gv
f
g
+ v
ov
ρ
V
f
o
= 0, (3.38)
ϕ
t
ρ
D
(1 v
ov
)s
o
+
x
ρ
D
(1 v
ov
)uf
o
= 0, (3.39)
t
H
r
+
ϕ
s
o
(
v
ov
ρ
V
h
oV
+
ρ
D
(1 v
ov
)h
oD
)
+
ϕ
s
g
ρ
gw
h
gW
+
ρ
gv
h
gV

+
+
x
u
f
o
(
v
ov
ρ
V
h
oV
+
ρ
D
(1 v
ov
)h
oD
)
+
f
g
ρ
gw
h
gW
+
ρ
gv
h
gV

= 0,
(3.40)
where
ρ
gv
= y
gv
ρ
gV
,
ρ
gw
=(1 y
gv
)
ρ
gW
and y
gv
depend on T and v
ov
.
A particular case occurs in the absence of dead oil. Then v
ov
= 1 trivially and Eq.
(3.39) disappears.
Notice for this physical situation i
I
(s
w
,s
g
, T,v
ov
, y
gv
) :=(s
g
, T,v
ov
); the terms R
T
G
and R
T
F
are functions with 6 components given by:
R
T
G
=
0,
ϕρ
gw
s
g
,
ϕρ
gv
s
g
,
ϕρ
V
v
ov
s
o
,
ϕρ
D
(1 v
ov
)s
o
,
H
r
+
ϕ
s
o
(
ρ
V
v
ov
h
oV
+
ρ
D
(1 v
ov
)h
oD
)
+
ϕ
s
g
ρ
gw
h
gW
+
ρ
gv
h
gV

.
and
R
T
F
=
0,
ρ
gw
f
g
,
ρ
gv
f
g
,
ρ
V
v
ov
f
o
,
ρ
D
(1 v
ov
) f
o
,
f
o
(
ρ
V
v
ov
h
oV
+
ρ
D
(1 v
ov
)h
oD
)
+
f
g
ρ
gw
h
gW
+
ρ
gv
h
gV

.
58
Liquid oleic and aqueous phases.
There are three components. Using the Gibb’s phase rule
(3.2), in addition to the
temperature T, we have another thermodynamical degree of freedom; we choose v
ov
to be this variable. In this physical situation y
gv
is meaningless, but if it is necessary,
it can be obtained for continuity in phase space as function of T and v
ov
using Eq.
(3.12).
The primary variables to be determined are s
w
, T, v
ov
,soP
IV
= {s
w
, T,v
ov
}; the
speed u is not constant; and the trivial variables are s
g
= 0 and y
gv
,soP
IV
= {s
g
=
0, y
gv
}. The physical situation is characterized by
IV
= {s
w
,s
g
= 0,T,v
ov
, y
gv
}.
The system
(3.29), (3.30)(3.27), (3.28) reduces to the following system of conser-
vation for water, volatile oil, dead oil and energy:
ϕ
t
ρ
W
s
w
+
x
u
ρ
W
f
w
= 0, (3.41)
ϕ
t
v
ov
ρ
V
s
o
+
x
uv
ov
ρ
V
f
o
= 0, (3.42)
ϕ
t
ρ
D
(1 v
ov
)s
o
+
x
ρ
D
(1 v
ov
)uf
o
= 0, (3.43)
t
H
r
+
ϕ
s
w
ρ
W
h
W
+
ϕ
s
o
(
v
ov
ρ
V
h
oV
+
ρ
D
(1 v
ov
)h
oD
)
+
+
x
u
f
w
ρ
W
h
W
+ f
o
(
v
ov
ρ
V
h
oV
+
ρ
D
(1 v
ov
)h
oD
)
= 0. (3.44)
Two particular cases arise: one of them corresponds to absence of dead oil, the other
to the absence of volatile oil.
Liquid water phase and gaseous steam / volatile oil phase.
There are two components, in the absence of dead oil. Using Gibb’s phase rule
(3.2),
we see that there are two thermodynamical degrees of freedom. Since the pressure is
fixed, the only degree of freedom is the temperature. Since there is no liquid volatile
oil v
ov
is not defined. On the other hand, since P
v
(T)=P P
w
(T), y
gv
is uniquely
defined by the temperature T, see Sec. 3.2.1.
The primary variables to be determined are s
w
, s
g
, T,soP
V
= {s
w
,s
g
, T}; the speed
u is not constant; and the trivial variables are v
ov
= 0 and y
gv
,soP
V
= {v
ov
, y
gv
}.
The physical situation is characterized by
V
= {s
w
,s
g
, T,v
ov
= 0, y
gv
}.
The system
(3.29), (3.30)(3.27), (3.28) reduces to the following system of conser-
59
vation laws for water, volatile oil and energy:
ϕ
t
ρ
W
s
w
+
x
u
ρ
W
f
w
+ f
g
ρ
gw
= 0, (3.45)
ϕ
t
ρ
gv
s
g
+
x
u
ρ
gv
f
g
= 0, (3.46)
t
H
r
+
ϕ
s
w
ρ
W
h
W
+
ϕ
s
g
ρ
gw
h
gW
+
ρ
gv
h
gV

+
x
u
f
w
ρ
W
h
W
+ f
g
ρ
gw
h
gW
+
ρ
gv
h
gV

= 0.
(3.47)
The steam fraction in the gaseous phase is always positive. One particular case arises,
corresponding to the absence of volatile oil.
3.2.3 One phase
In the following subsections, the physical situations consist of a single phase, so p = 1
in the Gibb’s phase rule
(3.2).
Liquid oleic phase.
There are two components. Using the Gibb’s phase rule
(3.2) we can see that in
addition to the temperature T there is another thermodynamical degree of freedom.
The suitable thermodynamical variable is v
ov
. As in the previous physical situation,
y
gv
is meaningless, but it can be obtained from continuity in phase space as a function
of T and v
ov
using Eq. (3.12).
The primary variables to be determined are T, v
ov
,soP
VI
= {T, v
ov
}; the speed u
is not constant; the trivial variables are s
w
= 0, s
g
= 0 and y
gv
,soP
VI
= {s
w
= 0,s
g
=
0, y
gv
}. The physical situation is characterized by
VI
= {s
w
= 0,s
g
= 0,T,v
ov
, y
gv
}.
The system of equations
(3.29), (3.30)(3.27), (3.28) reduces to the conservation
laws for volatile oil, dead oil and energy:
ϕ
t
v
ov
ρ
V
+
x
uv
ov
ρ
V
= 0, (3.48)
ϕ
t
(1 v
ov
)
ρ
D
+
x
u
(1 v
ov
)
ρ
D
= 0, (3.49)
t
H
r
+
ϕ
v
ov
ρ
V
h
oV
+(1 v
ov
)
ρ
D
h
oD
+
x
u
v
ov
ρ
V
h
oV
+(1 v
ov
)
ρ
D
h
oD
= 0.
(3.50)
Liquid aqueous phase.
There is one component. Using the Gibb’s phase rule
(3.2), the only thermodynamical
variable is the temperature T. Since there is no oil in this situation, v
ov
and y
gv
are
not defined in principle, but we obtain them as limits of Eqs.
(3.33) and (3.12).
Since
ρ
W
does not depend on T, one can show that u is constant spatially.
60
The primary variable to be determined by the system of equations is T,soP
VII
=
{
T}; the speed u is constant; and the trivial variables are s
w
= 1 , s
g
= 0 and v
ov
and
y
gv
so P
VII
= {s
w
= 1,s
g
= 0,v
ov
, y
gv
}. The physical situation is characterized by
VII
= {s
w
= 1,s
g
= 0,T,v
ov
, y
gv
}.
The system of equations
(3.29), (3.30), (3.27), (3.28) reduces to the single scalar
equation for conservation laws of energy:
t
H
r
+
ϕρ
W
h
W
+ u
x
ρ
W
h
W
= 0. (3.51)
Gaseous steam/volatile oil.
There are two components. Since there is no liquid oil v
ov
is not defined, but we
obtain it as a limit of Eq.
(3.33). Using the Gibb’s phase rule (3.2), in addition to the
temperature T, we have another thermodynamical degree of freedom; we choose y
gv
as this variable.
The primary variables to be determined are T, y
gv
,soP
VIII
= {T, y
gv
}; the speed
u is not constant; the trivial variables are s
w
= 0, s
g
= 1 and v
ov
, P
VIII
= {s
w
= 0,s
g
=
1,v
ov
}. The physical situation is characterized by
VIII
= {s
w
= 0,s
g
= 1,T,v
ov
, y
gv
}.
The system of equations governing this flow reduces to the conservation laws of
steam, of volatile oil in the gaseous phase and of energy:
ϕ
t
(1 y
gv
)
ρ
gW
+
x
u
(1 y
gv
)
ρ
gW
= 0, (3.52)
ϕ
t
y
gv
ρ
gV
+
x
uy
gv
ρ
gV
= 0, (3.53)
t
H
r
+
ϕ
(1 y
gv
)
ρ
gW
h
gW
+ y
gv
ρ
gV
h
gV

+
x
u
(1 y
gv
)
ρ
gW
h
gW
+ y
gv
ρ
gV
h
gV
= 0.
(3.54)
Pure steam phase.
In this case there are one phase and one component. Using the Gibb’s phase rule
(3.2), the only degree of freedom is the temperature. Since there is no oil, v
ov
and y
gv
are not defined, but they can be obtained as limits of Eqs. (3.33) and (3.12).
The primary variables to be determined is T,so
P
IX
= {T}; the speed u is not
constant; and the trivial variables are s
w
= 0, s
g
= 1, v
ov
and y
gv
,soP
IX
= {s
w
=
0,s
g
= 1,v
ov
, y
gv
}. The physical situation is characterized by
VIII
= {s
w
= 0,s
g
=
1,T,v
ov
, y
gv
}.
The system reduces to the conservation laws of steam and of energy:
ϕ
t
ρ
gw
+
x
u
ρ
gw
= 0, (3.55)
t
H
r
+
ϕρ
gw
h
gW
+
x
u
ρ
gw
h
gW
= 0. (3.56)
61
The next cases consists of single phase and single components, so in the Gibb’s
phase rule p
= 1 and c = 1, so there is only two thermodynamical degrees of freedom.
Since the pressure is fixed, the temperature T is chosen. This will be the primary
variable.
Pure liquid volatile oil.
Since there is only volatile oil in the oleic phase, v
ov
= 1. Since there is no gaseous
phase y
gv
is not defined, but they are obtained by continuity as a limit process.
The primary variable is T,so
P
X
= {T}. The speed u is constant; the trivial
variables are s
w
= 0, s
g
= 0, v
ov
= 1 and y
gv
,soP
X
= {s
w
= 0,s
g
= 0,v
ov
= 1, y
gv
}.
The physical situation is characterized by
X
= {s
w
= 0,s
g
= 0,T,v
ov
= 1, y
gv
}.
The system of equations reduces to the conservation laws of energy:
t
H
r
+
ϕρ
V
h
oV
+ u
x
ρ
V
h
oV
= 0. (3.57)
Pure gaseous volatile oil.
Since there is only volatile oil in the gaseous phase, thus y
gv
= 1; Since there is no
liquid oil v
ov
is not defined, but can be obtained by continuity as a limit process.
The primary variable is the T,so
P
XI
= {T}. The speed u is not constant; the
trivial variables are s
g
= 0, s
w
= 0, v
ov
and y
gv
= 1,soP
XI
= {s
w
= 0,s
g
= 0,v
ov
, y
gv
=
1}. The physical situation is characterized by
XI
= {s
w
= 0,s
g
= 0,T,v
ov
, y
gv
= 1}.
The system of equations reduces to the conservation laws of gaseous volatile oil
and of energy:
ϕ
t
ρ
gV
+
x
u
ρ
gV
= 0, (3.58)
t
H
r
+
ϕρ
gV
h
gV
+
x
u
ρ
gV
h
gV
= 0. (3.59)
Pure liquid dead oil.
Since the the vapor pressure is negligible in the dead oil, it occurs just in the oleic
phase and y
gv
is not defined, but it is obtained by continuity as a limit process.
The primary variable is T,so
P
XII
= {T}. The speed u is constant; the trivial
variables are s
w
= 0, s
g
= 0, v
ov
= 0 and y
gv
,soP
XI
= {s
w
= 0,s
g
= 0,v
ov
= 0, y
gv
}.
The physical situation is characterized by
XI
= {s
w
= 0,s
g
= 0,T,v
ov
= 0, y
gv
}.
The system of equations reduces to to the conservation laws of energy:
t
H
r
+
ϕρ
D
h
oD
+ u
x
ρ
D
h
oD
= 0. (3.60)
62
CHAPTER 4
The Riemann solution for the injection of steam and
nitrogen
The frequent and widespread occurrence of contamination due to spills and leaks
of organic materials, such as petroleum products, that occur during their transport,
storage and disposal constitute a menace to our high-quality groundwater resources.
In spite of increased awareness of the environmental impacts of oil spills, it appears
to be impossible to avoid these accidents, so it is necessary to develop techniques for
groundwater remediation.
Steam injection is widely studied in Petroleum Engineering, see [10]. In Chapter 5,
we study the steam and water injection in several proportions into a porous medium
saturated with a different mixture of steam and water. A disadvantage of pure steam
injection is the ecological impact of high temperatures. This can be alleviated if we co-
inject nitrogen leading to a lower temperatures. Within of such a mixture, the boiling
temperature of water depends on nitrogen concentration and it is lower than water
regular boiling temperature.
Another application is the recovery of geothermal energy. Injecting water and
nitrogen into a hot porous rock, the water in the mixture evaporates and rock thermal
heat can be recovered even if the rock is not very hot.
We present a physical model for steam, nitrogen and water flow based on mass
balance and energy conservation equations. We present the main physical definitions
and equations of the model. We study the three possible physical phase mixture
situations. For each physical situation, we reduce the original four balance equations
system to be presented in Section 4.2 to systems of conservation laws of type
(2.2):
t
G
(V)+
x
uF
(V)=0,
supplemented by appropriate thermodynamic constraints between variables, such as
Raoult’s law and others described in Appendix A. Here V
=(V
1
,V
2
) : R ×R
+
−→ D
63
R
3
is a subset of the variables: gas saturation s
g
, steam composition
ψ
gw
and temper-
ature T and it represents the unknowns in each physical situation; G
=(G
1
, G
2
, G
3
) :
D −→ R
3
and F =(F
1
, F
2
, F
3
) : D −→ R
3
are the accumulation vector and the flux
vector, respectively; u is a total velocity.
The state of the general system is represented by
(s
g
,
ψ
gw
, T,u). Systems of conser-
vation equations of the type
(2.2) have an important feature: the variable u does not
appear in the accumulation terms, it appears isolated in the flux terms, thus such sys-
tems have an infinite speed mode associated to u; despite this fact and the presence of
source terms in the original balance equations we are able to solve the associated Rie-
mann problem, which is a mathematical novelty. Another surprise in this model is the
presence of a rarefaction wave associated to evaporation in the two-phase situation.
In Section 4.1, we present the physical model that describes the injection of steam
and nitrogen in a one-dimensional horizontal porous rock initially filled with water
liquid. In Section 4.2, we present the model equations for balance of water, steam,
nitrogen and energy. The full governing equations are not a system of conservation
laws, since there is a water mass transfer term, which is the condensation or evapora-
tion of water between the liquid and gaseous phases. In Section 4.3, we describe three
physical situations in which the balance system is rewritten as conservation systems
of type
(2.2). In Section 4.4, we study the type of waves that appear in the solution
and some general results for 3
×3 systems; a more complete theory is established in
[2 ]. In Secs. 4.5, 4.6 and 4.7, we obtain the waves in each physical situation, it is
remarkable that there is an evaporation rarefaction wave in the two-phase situation.
In Section 4.8, we study the shocks between different regions. In Section 4.9, we show
an example of the Riemann solution for a mixture of steam and nitrogen in the spg
injected into a rock initially filled with water.
4.1 Physical model
We consider the injection of steam and nitrogen in a one-dimensional horizontal
porous rock core initially filled with water at temperature T
0
. The core consists
of the rock with constant porosity
ϕ
and absolute permeability k (see Appendix A).
We assume that the fluids are incompressible and that the pressure variations along
the core are so small that they do not affect the physical properties of the fluids.
Darcy’s law for multiphase flow relates the pressure gradient in each fluid phase
with its seepage speed:
u
w
=
kk
rw
µ
w
p
w
x
, u
g
=
kk
rg
µ
g
p
g
x
. (4.1)
The water and gas relative permeability functions k
rw
and k
rg
are considered to be
functions of their respective saturations (see Appendix A);
µ
w
and
µ
g
are the vis-
cosities of liquid and gaseous phases; p
w
and p
g
are the pressures in the liquid and
gaseous phases. The capillary pressure P
c
and the capillary diffusion coefficient
64
are:
P
c
= P
c
(s
w
)=p
g
p
w
and = f
w
kk
rg
µ
g
dP
c
ds
w
0. (4.2)
The fractional flow functions for water and steam are defined by:
f
w
=
k
rw
/
µ
w
k
rw
/
µ
w
+ k
rg
/
µ
g
, f
g
=
k
rg
/
µ
g
k
rw
/
µ
w
+ k
rg
/
µ
g
. (4.3)
Using Darcy’s law (3.14) and the definition of P
c
in Eq. (4.2.a), Eqs. (4.2) and (3.15)
yield:
u
w
= uf
w
s
w
x
, u
g
= uf
g
s
g
x
, where u
= u
w
+ u
g
(4.4)
is the total or Darcy velocity. We will see that acts as a capillary diffusion coeffi-
cient; s
w
is the water saturation, i.e., the pore fraction filled with water; similarly, s
g
is the gas saturation.
4.2 The model equations
Neglecting molecular diffusion effects, we write the equations of mass balance for
water, steam and nitrogen as:
t
ϕρ
W
s
w
+
x
u
w
ρ
W
=+q
g−→ a,w
, (4.5)
t
ϕρ
gw
s
g
+
x
u
gw
ρ
gw
= q
g−→ a,w
, (4.6)
t
ϕρ
gn
s
g
+
x
u
gn
ρ
gn
= 0, (4.7)
where q
g−→ a,w
is the water mass source term (i.e., the condensation rate between
the steam and water phases); here
ρ
W
is the water density, which is assumed to be
constant,
ρ
gw
(
ρ
gn
) denote the concentration of steam (nitrogen) in the gaseous phase
(mass per unit gas volume); the concentration dependence on temperature is specified
by Raoult’s laws and other thermodynamical relationships given in Appendix A. The
saturations s
w
and s
g
add to 1. By (4.3), the same is true for f
w
and f
g
.
Since there are 4 unknowns (T, s
g
,
ψ
gw
and u) and 3 equations, it is necessary
to specify another equation, representing energy conservation, based on an enthalpy
formulation, see
[2,3]. We include longitudinal heat conduction, but neglect heat
losses to the surrounding rock. We ignore adiabatic compression and decompression
effects. Thus the energy conservation is given by:
t
ϕ
ˆ
H
r
+
ρ
W
h
W
s
w
+(
ρ
gw
h
gW
+
ρ
gn
h
gN
)s
g
+
x
u
w
ρ
W
h
W
+ u
g
(
ρ
gw
h
gW
+
ρ
gn
h
gN
)
=
x
κ
T
x
, (4.8)
65
here H
r
is the rock enthalpy per unit volume and
ˆ
H
r
= H
r
/
ϕ
; h
W
, h
gW
and h
gN
are the
enthalpies per unit mass of water in the liquid aqueous phase, water and nitrogen
in the gaseous phase; these enthalpies depend on temperature as described by Eqs.
(A.3) and (A.4). The composite conductivity of the rock/fluid system
κ
>
κ
r
> 0 is
given by:
κ
=
κ
r
+
ϕ
(s
w
κ
w
+ s
g
κ
g
). (4.9)
We define H
W
and H
g
, the effective water and gas enthalpies per unit volume as:
H
W
=
ρ
W
h
W
and H
g
=
ρ
gw
h
gW
+
ρ
gn
h
gN
, (4.10)
so from
(4.4),(4.10), we can rewrite (4.5)-(4.8) as:
t
ϕρ
W
s
w
+
x
u
ρ
W
f
w
=
x
ρ
W
s
w
x
+ q
g−→ a,w
, (4.11)
t
ϕρ
gw
s
g
+
x
u
ρ
gw
f
g
=
x
ρ
gw
s
g
x
q
g−→ a,w
, (4.12)
t
ϕρ
gn
s
g
+
x
u
ρ
gn
f
g
=
x
ρ
gn
s
g
x
, (4.13)
t
ϕ
ˆ
H
r
+ H
w
s
w
+ H
g
s
g
+
x
u
H
w
f
w
+ H
g
f
g
=
=
x
H
g
H
w
s
g
x
+
x
κ
T
x
. (4.14)
We are interested in scales dictated by field reservoirs. The effect of spatial second
derivative terms (capillary pressure, heat conductivity, etc) is to widen the heat con-
densation front as well as other shocks, while the convergence of the characteristics
tries to sharpen them. The balance of these effects yields the width of these fronts.
In the field this width is typically a few tenth of centimeters; on the other hand, the
distance between injection and production wells is of the order of 1000 meters. Thus
this width is negligible, so we can set it to zero and simplify our analysis with no error
of practical importance. From now on we disregard the diffusive terms in (4.11)-(4.14),
resulting in the following balance system:
t
ϕρ
W
s
w
+
x
u
ρ
W
f
w
=+q
g−→ a,w
, (4.15)
t
ϕρ
gw
s
g
+
x
u
ρ
gw
f
g
= q
g−→ a,w
, (4.16)
t
ϕρ
gn
s
g
+
x
u
ρ
gn
f
g
= 0, (4.17)
t
ϕ
ˆ
H
r
+ H
w
s
w
+ H
g
s
g
+
x
u
H
w
f
w
+ H
g
f
g
= 0. (4.18)
4.3 Physical situations
There are three regions in different physical situations. Initially, there is only water
in the core; we call this situation the single-phase liquid situation, spl. We can inject a
66
superheated gas to form a region containing only gaseous steam and nitrogen; we call
this the single-phase gaseous situation, spg. There is also the two-phase situation,
tp, where there is a mixture of liquid water, gaseous nitrogen and steam at boiling
temperature. The latter depends on the concentration of nitrogen in the gas. We
assume that each situation is in local thermodynamical equilibrium, so we can use
Gibbs’ phase rule
(3.2).
4.3.1 Single-phase gaseous situation - spg
In this situation there are two components and one gaseous phase, i.e., c = 2 and
p
= 1, so from Gibbs’ phase rule (3.2) there are three thermodynamical degrees of
freedom. Since in our thermodynamical model the pressure is fixed, there are two
unknown thermodynamical variables: temperature and gas composition.
We assume that nitrogen and steam in the gaseous phase behave as ideal gases
with densities depending on T denoted by
ρ
gN
and
ρ
gW
, see (A.9) in Appendix A. We
use Raoult’s law (see
[68]) and assume that there are no volume effects due to mixing,
so that the volumes of the components are additive:
ρ
gw
/
ρ
gW
(T)+
ρ
gn
/
ρ
gN
(T)=1. (4.19)
This fact allows to define the steam and nitrogen gas composition, respectively, as:
ψ
gw
=
ρ
gw
/
ρ
gW
(T) and
ψ
gn
=
ρ
gn
/
ρ
gN
(T) , so
ψ
gw
+
ψ
gn
= 1, (4.20)
thus when the steam or nitrogen composition is known, the other composition is ob-
tained trivially from (4.20.c). We use
ψ
gw
as the basic state variable. The composi-
tions
ρ
gw
and
ρ
gn
are functions of temperature and composition; they are obtained
from
(4.20.a) and (4.20.b).
Therefore, there are three unknowns to be determined: temperature T, gas com-
position
ψ
gw
and speed u. The saturations are trivial, s
g
= 1 and s
w
= 0.
We rewrite equations (4.15)-(4.18) using Eqs.
(4.20.a) and (4.20.b). Since s
w
=
0, f
w
= 0, so Eq. (4.15) disappears and we obtain that q
g−→ a,w
vanishes; thus the
remaining system becomes:
t
ϕθ
W
ψ
gw
T
1
+
x
u
θ
W
ψ
gw
T
1
= 0, (4.21)
t
ϕθ
N
ψ
gn
T
1
+
x
u
θ
N
ψ
gn
T
1
= 0, (4.22)
t
ϕ
ˆ
H
r
+
ψ
gw
H
gW
+
ψ
gn
H
gN
+
x
u
ψ
gw
H
gW
+
ψ
gn
H
gN
= 0; (4.23)
we have substituted
ρ
gW
and
ρ
gN
from Eq. (A.9), and defined
θ
W
,
θ
N
, H
gW
(T) and
H
gN
(T):
θ
W
=
M
W
p
at
R
,
θ
N
=
M
N
p
at
R
, H
gW
(T)=
θ
W
h
gW
T
and H
gN
(T)=
θ
N
h
gN
T
, (4.24)
67
where M
N
, M
W
are the molar masses of nitrogen and water; R is the universal gas
constant; p
at
is the atmospheric pressure; h
gW
and h
gN
are functions of T given in
Appendix A.
Remark 4.3.1. The spg does not exist for any pair
(T,
ψ
gw
), because of thermodynamic
constraints. The tp situation occurs for states where the composition is given in terms
of the temperature T as:
ψ
gw
(T)=
ρ
gw
(T)/
ρ
gW
(T) , (4.25)
where
ρ
gw
and
ρ
gW
given by Eqs. (A.8. a), (A.9. a). The composition
ψ
gw
(T) is the pro-
jection of the tp on the plane
{T,
ψ
gw
}. As we assume thermodynamical equilibrium,
the solution of Eq.
(4.25) defines the boundary of the physical region in the plane
{T,
ψ
gw
}, see Fig. (4.1. a).
We define the spg physical region as:
Γ
=
(T,
ψ
gw
) spg that satisfy
ψ
gw
ρ
gw
(T)/
ρ
gW
(T)
. (4.26)
For each
ψ
gw
, the temperature that satisfies (4.25) is called saturation temperature.
The curve Γ is formed by the pairs
(T,
ψ
gw
(T)) that satisfy (4.25).
260 280 300 320 340 360 380 400 420
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Temperature /K
Steam Composition
Physical Region
Water boiling temperature
Γ
Steam region
Boiling region
Steam regionSteam region
Water region
Figure 4.1: a) Left: The physical region lies to the right of the curve defined by (4.25).
The continuous graph represents the saturation temperature for the mixture of water
and nitrogen. b) Right: Phase space for V
=(s
W
,
ψ
gw
, T) and physical situations. The
bold lines represent the physical regions for water and steam without nitrogen, see
Chapter 5.
4.3.2 Two-phase situation - tp
In the tp there are two components, c = 2, and two-phases, p = 2 ; using Gibbs’
phase rule
(3.2), there are two thermodynamical degrees of freedom, temperature
and pressure (which is fixed). The compositions in Eqs.
(4.27) and (4.28) depend on
temperature, and vice-versa inverting Eq.
(4.25). As the pressure is given, the boiling
68
temperature of water is specified by these compositions. When pure steam is injected
the condensation temperature is constant T
b
= 373.15, see Appendix A. We have three
variables to be determined: temperature, saturation and total Darcy velocity.
There is a linearly dependent equation in the system (4.15)-(4.18). We add (4.15)
and (4.16) and replace (4.15)-(4.18) by:
t
ϕ
ρ
W
s
w
+
ρ
gw
s
g
+
x
u
ρ
W
f
w
+
ρ
gw
f
g
= 0, (4.27)
t
ϕ
ρ
gn
s
g
+
x
u
(
ρ
gn
f
g
)=0, (4.28)
t
ϕ
ˆ
H
r
+ H
W
s
w
+ H
g
s
g
+
x
u
(H
W
f
w
+ H
g
f
g
)=0, (4.29)
The compositions
ρ
gw
and
ρ
gn
are functions of temperature only that can be obtained
from Eqs.
(A.8. a) and (A.8.b) in Appendix A.
4.3.3 Single-phase liquid situation - spl
In the spl there is one component and one phase, so from Gibbs’ phase rule (3.2),
there are two thermodynamical degrees of freedom, temperature and pressure (which
is fixed). Since s
w
= 1 and s
g
= 0,wehavef
w
= 1 and f
g
= 0. From (4.12), q
g−→ a,w
vanishes. One can prove that the total Darcy velocity u is independent of position, so
(4.15)-( 4.18) become:
t
ϕ
ˆ
H
r
+ H
w
+ u
W
H
w
x
= 0, (4.30)
where we use u
W
to indicate that the velocity is spatially constant in the spl; V is
just the temperature T. We assume that rock and water enthalpy depend linearly on
temperature, so we can rewrite
(4.30) as:
t
T
+
λ
W
T
T
x
= 0, where
λ
W
T
=
u
W
ϕ
C
W
C
W
+
ˆ
C
r
, (4.31)
and C
W
is the water rock heat capacity and
ˆ
C
r
is the rock heat capacity C
r
divided by
ϕ
. All quantities are given in Appendix A.
4.3.4 Primary, secondary and trivial variables
Fig. 4.1.b shows the 3 physical situations in the variables V =(T,
ψ
gw
,s
w
). As basic
variable, we choose s
w
instead of s
g
for convenience.
For each physical situation there are three groups of variables. One group en-
compasses the variables V in the system
(2.2). We call the variables in V “primary
variables” or basic variables. The variable u is called “secondary variable” because we
will see that it is obtained from the primary variables. The trivial variables are the
variables that depend on primary variables in a simple way (for example, in the tp the
steam composition depends on temperature) or are obtained trivially (for example, in
69
the spg the gas saturation is trivial and its value is 1). In Table 4.3.4 we summarize
these variables.
Physical situation \ Variable types Primary Secondary Trivial
Single-phase gaseous situation T,
ψ
gw
u s
g
= 1
Two-phase situation s
g
, T u
ψ
gw
(T)
Single-phase liquid situation T u s
g
= 0,
ψ
gw
Table 1: Classification according to situation.
4.4 General theory of Riemann Solutions
We are interested in the Riemann problem associated to (4.15)-(4.18) , that is the so-
lution of these equations with initial data
(s
g
,
ψ
gw
, T,u)
L
if x > 0
(s
g
,
ψ
gw
, T,·)
R
if x < 0,
(4.32)
The speed u
L
> 0 is specified at the injection point. In the next sections we show that
u
R
can be obtained as function of u
L
and the primary variables.
The general solution of the Riemann problem associated to Eq.
(2.2) consists of
a sequence of elementary waves, rarefactions and shocks; they are studied in the
sections that follow,
[60]. We assume that the cumulative and flux functions have
continuous second derivative (
C
2
) within each physical situation and that are only
continuous between the physical situations.
4.4.1 Characteristic speeds in each physical situation
System of conservation laws in different forms must be used to find the characteristic
speeds. If we assume that the solution is sufficiently smooth, we differentiate all
equations in ( 2.2) with respect to their variables, obtaining a system of the form 2.27
)
B
t
V
u
+ A
x
V
u
= 0, (4.33)
where the matrices B and A (which depend on V) are the derivatives of G
(V) and
uF
(V) with respect to the variables V and u. Since G(V) does not depend on u, the
last column in the matrix B is zero. The characteristic values
λ
i
:=
λ
i
(V,u) and vectors
r
i
:=
r
i
(V,u), (where i is the label of each eigenvector) for the following system are the
rarefaction wave speeds and directions:
A
r
i
=
λ
i
B
r
i
where
λ
i
is obtained by solving det(A
λ
i
B)=0. (4.34)
Similarly the left eigenvectors
i
=(
i
1
,
i
2
,
i
3
) satisfy:
i
A =
λ
i
i
B, for the same
λ
i
. (4.35)
70
Remark 4.4.1. We consider here a general 3 × 3 system because it is the simplest
non-trivial system of type
(2.2). Generically, the right and left eigenvectors have three
components: two primary variables and one secondary variable. However, in the spl
situation we do not utilize this formalism because the system
(2.2) reduces to a single
equation.
Hereafter , the word “eigenvectors” represents “right eigenvectors”. For brevity,
we will write the right and left eigenvectors without the upper arrow, i.e., we replace
r by r and
by .
Remark 4.4.2. In Lemma
[2.2.1], we prove that the left eigenvectors
i
do not depend
on the Darcy speed u.
Definition 4.4.1. For each i, the integral curves in the
(V,u) plane are solutions of
dV
d
ξ
,
du
d
ξ
= r, i.e,
dV
1
d
ξ
= r
1
,
dV
2
d
ξ
= r
2
and
du
d
ξ
= r
3
. (4.36)
When
ξ
satisfies:
ξ
=
λ
(V(
ξ
),u(
ξ
)), (4.37)
the integral curves define rarefaction curves. If in addition to
(4.37), the variable
ξ
increases and satisfies x =
ξ
t the integral curves define rarefaction waves in the (x,t)
plane.
Remark 4.4.3. In Secs. 4.5 and 4.7, we calculate the eigenvalues and eigenvectors for
the nitrogen-steam-water model and show that they have the following form:
λ
= u(
ξ
)
ˆ
λ
(V(
ξ
)) and r =
(
η
1
(V(
ξ
)),
η
2
(V(
ξ
)),u(
ξ
)
η
3
(V(
ξ
)
)
, (4.38)
where
ˆ
λ
and
η
j
for j = 1,2,3 do not depend on u. It is useful to define:
ˆ
λ
i,
(V) :=
λ
i
/u
,
ˆ
λ
i,+
(V) :=
λ
i
/u
+
and
ˆ
r
i
=(r
i
1
,r
i
2
) for i = 1,2, (4.39)
where u
is the Darcy speed for V
and u
+
is the Darcy speed for V
+
. The primary
variables V
and V
+
are the first and last value of V in the rarefaction wave. Then we
can see from Eq.
(4.36) that (V, u) along the rarefaction wave satisfy:
dV
1
d
ξ
,
dV
2
d
ξ
=
η
1
(V(
ξ
)),
η
2
(V(
ξ
))
and
du
d
ξ
= u
η
3
(V(
ξ
)). (4.40)
Since V
=(V
1
,V
2
) does not depend on u we can solve first (4.40. a) obtaining V(
ξ
) and
then substitute it in
(4.40.b) to obtain u(
ξ
):
u
(
ξ
)=u
ex p
ξ
ξ
η
3
(V(
σ
))d
σ
. (4.41)
71
Here u
is the first value for u on this integral curve, and
ξ
is chosen such that
ξ
=
λ
(V
,u
). Notice that
ξ
is given implicitly by (4.37); using (4.41) this implicit
relation is:
ξ
= u
ˆ
λ
(V(
ξ
))ex p
ξ
ξ
η
3
(V(
σ
))d
σ
. (4.42)
ξ
depends on u
and V, but it does not depend on the Darcy speed in the rarefaction
curve. When V
= V
+
, i.e., the last V in the rarefaction wave, we denote
ξ
=
ξ
+
.
Remark 4.4.4. Assume that u
> 0, so we perform the change of variable:
ξ
=
x
t
−→
ˆ
ξ
=
x
u
t
, (4.43)
The system
(4.33) can be written in terms of the independent variables (x, T)=(x,u
t).
In this space, one can prove that the eigenvalues and eigenvectors have the form
λ
i,
(V);
r
i
1
and r
i
2
are the first two components of the vector r
i
in Eq. (4.38.b). From Eqs. (4.37)
and (4.42), it follows that
ˆ
ξ
and
ˆ
ξ
satisfy:
ˆ
ξ
=
ˆ
λ
(V
), and
ˆ
ξ
=
ˆ
λ
(V(
ˆ
ξ
))ex p
ˆ
ξ
ˆ
ξ
η
3
(V(
σ
))d
σ
. (4.44)
It follows that in
(x, T) the variables V and
ˆ
ξ
do not depend on u, so the rarefaction
curves too do not depend on u . The Darcy speed keeps the form
(4.41).
4.4.2 Shock waves
Frequently, discontinuities or shocks appear in solution of the Riemann (or Cauchy)
problems of non-linear hyperbolic or hyperbolic-elliptic equations (see
[60]). The Rankine-
Hugoniot condition for shocks (RH) needs to be satisfied.
The main feature of this class of equations is the existence of separated regions
where the system of balance equations (4.15)-(4.18) reduces to different systems of
conservation equations
(2.2). In the nitrogen/steam/water flow problem considered in
this paper, there are three different situations: the spg, the tp and the spl. To obtain
the complete Riemann problem solution it is necessary to link these regions. As the
water mass source term acts in the infinitesimal space between physical situations,
we propose the existence of shocks linking these different regions. Thus we divide
the study of the shocks into different groups: shocks within a physical situation and
shocks between different physical situations.
For a fixed W
=(V
,u
) the Rankine-Hugoniot Locus, RH locus, parametrizes
the discontinuous solutions of Eq.
(4.15)-( 4.18), ie., it consists of the W
+
=(V
+
,u
+
)
that satisfy the following RH condition and it is denoted by RH(W
), see [2.3.1]:
v
s
G
+
(V
+
) G
(V
)
= u
+
F
+
(V
+
) u
F
(V
), (4.45)
where
(V
+
,u
+
) is the state at the right of the shock and (V
,u
) is the state at the
left of the shock; v
s
is the shock speed; G
+
( G
) and F
+
(F
) are the accumulation and
72
flux terms at the right (left) of the shock, which in general have different expressions.
We specify the state
(V
,u
) on the left hand side, but at the right u
+
is not specified,
it is obtained from the RH condition
(4.45). From the continuity of F and G at the
interface between different physical situations is clear that W
+
= W
is always a
solution of
(4.45), i.e., the () state always belongs to a branch of RH locus.
We define the function
H = H(W
;W
+
) as:
H := v
s
G
+
(V
+
) G
(V
)
u
+
F
+
(V
+
)+u
F
(V
), (4.46)
notice that the RH locus of W
are the states W that satisfy H( W
;W
+
) := 0.
Since in this work we are interested in connected branches (i.e., that contain the
() state), we use the following criterion for admissibility of shocks instead of the
viscosity profile criterion:
Definition 4.4.2. (T.P. Liu.
[47,48]) We call shock curve the parts V of Rankine-
Hugoniot curve where the shock speed decreases when V moves away from V
. When
we consider the waves in
(x, t) also, each point of the shock curve represents shock
wave. The shock curve parametrizes the
(+) states of admissible shocks waves (),
(+).
Rankine-Hugoniot locus
For a fixed W
=(V
,u
), we obtain the RH locus, or RH(W
), from Eq. (4.45),
which can be written as:
v
s
[G
1
]=u
+
F
+
1
u
F
1
, (4.47)
v
s
[G
2
]=u
+
F
+
2
u
F
2
, (4.48)
v
s
[G
3
]=u
+
F
+
3
u
F
3
, (4.49)
where
[G
i
]=G
+
i
G
i
, G
±
i
= G
±
i
(V
±
) and F
±
i
= F
±
i
(V
±
) for i = 1,2,3. We rewrite the
system
(4.47)-( 4.49) as:
[G
1
] F
+
1
F
1
[
G
2
]
F
+
2
F
2
[
G
3
]
F
+
3
F
3
v
s
u
+
u
= 0, which requires det
[G
1
] F
+
1
F
1
[
G
2
]
F
+
2
F
2
[
G
3
]
F
+
3
F
3
= 0
(4.50)
to have a non-trivial solution. Eq.
(4.50.b) yields the implicit expression H
e
= 0 with:
H
e
:=[G
1
]( F
+
3
F
2
F
3
F
+
2
)+[G
2
]( F
+
1
F
3
F
1
F
+
3
)+[G
3
]( F
+
2
F
1
F
2
F
+
1
). (4.51)
Since we can obtain the RH locus in V, it more useful denote it by
RH(V
) instead
of
RH(W
). Generically, the expression H
e
(V
;V
+
)=0 defines a 1-dimensional
structure, see Section 4.4.2. Notice that there are only 2 primary variables in V
+
for
the spg and the tp (see the following Section for definition). In the spl there is only
one scalar equation with the temperature as the primary variable, so
RH(T
) is the
whole temperature physical range.
73
Solving the system (4.50), we obtain v
s
and u
+
as functions of V
, V
+
and u
:
v
s
= u
F
+
2
F
1
F
+
1
F
2
F
+
1
[G
2
] F
+
2
[G
1
]
=
u
F
+
3
F
2
F
+
2
F
3
F
+
2
[G
3
] F
+
3
[G
+
2
]
=
u
F
+
1
F
3
F
+
3
F
1
F
+
3
[G
1
] F
+
1
[G
3
]
, (4.52)
u
+
= u
F
1
[G
2
] F
2
[G
1
]
F
+
1
[G
2
] F
+
2
[G
1
]
=
u
F
2
[G
3
] F
3
[G
2
]
F
+
2
[G
3
] F
+
3
[G
2
]
=
u
F
3
[G
1
] F
1
[G
3
]
F
+
3
[G
1
] F
+
1
[G
3
]
. (4.53)
Of course, Eqs.
(4.52) and (4.53) are valid if each denominator is non-zero.
Remark 4.4.5. In the definitions that follow, all wave structures can be obtained in the
space of primary variables V. Using u
+
= u
+
(V
,u
;V
+
) and v
s
:= v
s
(V
,u
;V
+
),
we define:
u
+
u
=
χ
(V
;V
+
),
ˆ
v
(V
;V
+
) :=
v
s
u
and
ˆ
v
+
(V
;V
+
) :=
v
s
u
χ
(V
;V+)
. (4.54)
We define the set of unordered index pairs for Eqs.
(4.52) and (4.53) as:
P = {{1,2}, {3,1}, {2,3}}. (4.55)
From this Section 4.4.2 and Rems. 4.4.3 and 4.4.4, we have proved the following:
Proposition 4.4.1. Let u
L
be positive. The primary variables V in the shock and
rarefaction curves do not depend on the left Darcy speed u
L
> 0. If a sequence of waves
and states solve the Riemann problem in the primary variables, for a given u
L
> 0 (or
u
R
> 0), then it is also a solution for any other u
L
> 0 (or u
R
> 0):
(V
L
,·) if x < 0
(V
R
,·) if x > 0.
(4.56)
Moreover, assume that for each m there is
{p, q}∈P such that the following inequality
is satisfied: F
+
q,m
[G
p,m
] F
+
p,m
[G
q,m
] = 0. Then u
R
is given by:
u
R
= u
L
ρ
1
l=1
ex p
ξ
+,l
ξ
,l
η
l
3
(V(
σ
))d
σ
ρ
2
m=1
F
q,m
[G
p,m
] F
p,m
[G
q,m
]
F
+
q,m
[G
p,m
] F
+
p,m
[G
q,m
]
, (4.57)
where
η
l
3
is the third component of eigenvector r,
ξ
,l
and
ξ
+,l
are the first and the last
values of
ξ
l
associated to the l-th rarefaction wave.
In the Proposition above, G
j,m
and F
j,m
represent the j-th components ( j = {1,2,3})
of G and F on the m-th shock wave. Similarly ex p
ξ
+,l
ξ
,l
η
l
3
(V(
σ
))d
σ
is computed
along the l-rarefaction curve.
A very important result follows from the Proposition 2.3.2 is:
Corollary 4.4.1. Assume the same hypothesis of Proposition 2.3.2, then the Riemann
solution can be obtained in each physical situation first in the primary variables V.
Then the Darcy speed can be obtained at any point of the space
(V,u) in terms of V
and u
L
by an equation analogous to Eq. (4.57).
74
Remark 4.4.6. Corollary 4.4.1 states that the secondary variable u can be obtained
from the primary variables, so the former usually will not appear in figures.
Proposition 4.4.2. (Quadruple Shock Rule.) In the tpand the spgsituations, consider
four primary points and three Darcy speeds, determining four states:
(V
,u
) in the
tp,
(V
+
,u
+
) in the spg and (V
M
,u
M
) and (V
,u
) both in the tp or the spg. Assume
that there are shocks between the following pairs of states:
(i)
(V
,u
) and (V
+
,u
+
) with speed v
,+
,
(ii)
(V
,u
) and (V
M
,u
M
) with speed v
,M
,
(iii)
(V
M
,u
M
) and (V
,u
) with speed v
M,
,
such that two speeds coincide, i.e., at least one of the following equalities is satisfied:
either v
,+
= v
,M
or v
,+
= v
M,
or v
,M
= v
,+
. (4.58)
If the following conditions
(I) to (III) are satisfied:
(I) G
(V
+
) G(V
) and G(V
) G(V
M
) are LI;
(II) V
+
and V
have one coordinate V
i
with coinciding values;
(III)
H
e
/V
j
= 0 for j = i for all V ∈RH(V
),
then
(1) V
=V
+
;
(2) u
= u
+
;
(3) all the three speeds are equal:
v
,+
= v
,M
= v
M,
. (4.59)
Proof: Disregarding the index in accumulation and flux terms, the RH conditions
for
(V
,u
)-(V
+
,u
+
), (V
,u
)-(V
M
,u
M
) and (V
M
,u
M
)-(V
,u
) are respectively:
v
,+
(G(V
+
) G(V
)) = u
+
F(V
+
) u
F(V
), (4.60)
v
,M
(G(V
M
) G(V
)) = u
M
F(V
M
) u
F(V
), (4.61)
v
M,
(G(V
) G(V
M
)) = u
F(V
) u
M
F(V
M
). (4.62)
Assume that now Eq.
(4.58.a) is satisfied. Substituting v
,+
= v
,M
= v in Eqs.
(4.60
) and (4.61) and subtracting Eq. (4.60) from (4.61), we obtain:
v
(G(V
+
) G(V
M
)) = u
+
F(V
+
) u
M
F(V
M
). (4.63)
Notice that Eqs.
(4.63) and (4.62) define implicitly the RH locus by H
e
(V
M
;V
+
)=
0 in the variables V
M
and V
+
. Since the RH locus depends solely on V
M
and the
accumulation and flux functions, we obtain that both RH locus defined by Eqs.
(4.63)
and (4.62) coincide. Since (I) to (III) are satisfied, then V
= V
+
.
Now from Eq.
(4.53), we notice for a fixed u
that the Darcy and shock speeds
depend solely on V
and V
+
. From Eqs. (4.62) and (4.63), we can see that the ()
and (+) states are the same for each expression and that they define the same RH
locus, so u
= u
+
and Eq. (4.59) is satisfied.
The other cases are proved similarly.
75
Degeneracies of RH Locus
We utilize the notation:
D
ij
(V
,V
+
)=F
+
i
[G
j
] F
+
j
[G
i
]. (4.64)
Proposition 4.4.1 is valid if the denominator of Eq.
(4.57) is non zero for some {i, j}∈
P. So it is necessary to study the behavior of the solution when D
ij
(V
,V
+
)=0 for
all
{i, j}∈P. For a fixed pair {i, j}∈P, it easy to prove that:
Lemma 4.4.1. Let
{V
,V
+
} satisfy H
e
(V
;V
+
)=0, where H
e
is given by Eq. (4.51).
If
D
ij
(V
,V
+
)=0, then one of following conditions is satisfied:
(i) F
+
i
F
j
F
+
i
F
j
= 0, (4.65)
or
(ii ) D
12
(V
,V
+
)=D
31
(V
,V
+
)=D
23
(V
,V
+
)=0. (4.66)
From this Lemma it follows immediately that:
Corollary 4.4.2. Let
{V
,V
+
} satisfy H
e
(V
;V
+
)=0.IfD
ij
(V
,V
+
) vanishes for
two index pairs
{i, j}∈P, then it vanishes for all pairs.
Proposition 4.4.3. If
D
ij
(V
,V
+
)=0 for all {i, j}∈P and (F
+
1
, F
+
2
, F
+
3
) = 0,we
obtain:
[G
k
]=
ρ
1
F
+
k
and F
k
=
ρ
2
F
+
k
, for k = 1,2,3. (4.67)
where
ρ
1
and
ρ
2
are constants depending on [G], F
and F
+
. Moreover, for
χ
defined
in
(4.54.a), the shock speed v
s
satisfies:
v
s
= u
χ
(V
;V
+
)
ρ
2
ρ
1
. (4.68)
Proof: Let u
> 0. Since D
ij
(V
,V
+
)=0 ∀{i, j}∈P, it follows that:
(F
+
2
[G
3
] F
+
3
[G
2
])e
1
+(F
+
3
[G
1
] F
+
1
[G
3
])e
2
+(F
+
1
[G
2
] F
+
2
[G
1
])e
3
= 0, (4.69)
where e
i
for i = 1,2,3 is the canonical basis for R
3
. Eq. (4.69) can be written as:
(F
+
1
, F
+
2
, F
+
3
) ×([G
1
],[G
2
],[G
3
]) = 0, (4.70)
where
× represents the curl. Since Eq. (4.70) is satisfied, it follows that [G] is parallel
to F
+
, so there is a constant
ρ
1
so Eq. (4.67. a) is satisfied. Substituting [G]=
ρ
1
F
+
into the RH condition (4.45), we obtain:
v
s
ρ
1
F
+
1
= u
+
F
+
1
u
F
1
, (4.71)
v
s
ρ
1
F
+
2
= u
+
F
+
2
u
F
2
, (4.72)
v
s
ρ
1
F
+
3
= u
+
F
+
3
u
F
3
, (4.73)
76
If F
i
= 0 for some i = 1,2,3, it follows that:
v
s
= u
+
/
ρ
1
. (4.74)
If F
i
= 0 for all i = 1,2,3, multiplying Eq. (4.71) by F
2
and (4.72) by F
1
and adding
the results, we obtain:
v
s
ϕρ
1
(F
+
1
F
2
F
+
2
F
1
)=u
+
(F
+
1
F
2
F
+
2
F
1
). (4.75)
Let us assume temporarily that F
+
1
F
2
F
+
2
F
1
= 0, so Eq. (4.74) is satisfied. Sub-
stituting v
s
given by Eq. (4.74) into Eq. (4.71) we obtain u
F
1
= 0. Since F
1
= 0,
generically, it follows that u
= 0, which is false. So F
+
1
F
2
F
+
2
F
1
= 0.
Similar calculations for Eqs.
(4.72), (4.73) and (4.71), (4.73) show that:
F
+
1
F
2
F
+
2
F
1
= F
+
2
F
3
F
+
3
F
2
= F
+
1
F
3
F
+
3
F
1
= 0. (4.76)
Since Eq.
(4.76) is satisfied, there exists a constant
ρ
2
such that F
=
ρ
2
F
+
. Eq.
(4.68) can be obtained by substituting [G]=
ρ
1
F
+
and F
=
ρ
2
F
+
in the RH condition
(4.45).
Remark 4.4.7. If (F
+
1
, F
+
2
, F
+
3
)=0 and ([G
1
],[G
2
],[G
3
]) = 0, it is easy to prove that:
F
k
=
ρ
3
[G
k
] and v
s
= u
ρ
3
, (4.77)
where
ρ
3
is a constant that depends on [G] and F
.
Corollary 4.4.3. The states
{V
,V
+
} satisfying the RH condition (4.45), for which
D
ij
(V
,V
+
)=0 for all {i, j}∈P, satisfy also:
F
+
i
F
j
F
+
j
F
i
= 0, ∀{i, j}∈P. (4.78)
We notice that the system
(4.78) has always the trivial solution V
+
= V
.
4.4.3 Bifurcation loci
Assume that there are no degeneracies any RH locus. For standard conservation laws,
there are loci where the solutions change topology such as: secondary bifurcation,
coincidence, double contact, inflection, hysteresis and interior boundary contact. The
proof of this change depends on the Bethe-Wendroff Theorem 2.3.6, that in this model
can be written as:
Proposition 4.4.4. Assume that F and G are
C
2
. Let v
s
(W
+
;W
) be the shock speed
between different physical situations. Assume that
i
(V
+
) ·(G
+
(V
+
) G
(V
)) = 0.
Then v
s
has a critical point at W
+
(and
ˆ
v
+
(V
;V
+
) has a critical point at V
+
), if and
only if:
ˆ
v
+
(V
;V
+
)=
λ
i,+
(V
+
) for i = 1 or 2, (4.79)
where
λ
+
(V) is given by Eq. (4.39) and v
+
(V
;V
+
) is given by (4.54).
77
Secondary bifurcation manifold
The RH locus for a fixed V
is obtained implicitly by H
e
(V
;V
+
)=0, where H
e
:
R
4
−→ R, and V
+
are the primary variables. At some points, these implicit expression
fail to define a curve for V
+
in the space of primary variables. We call the set of
these points the secondary bifurcation locus. From the implicit function theorem,
they are the
+ states for which there exists a state such that the following
equalities are satisfied:
H
e
(V
;V
+
)=0 and
H
e
V
+
j
= 0, for j = 1 and 2, (4.80)
where
H
e
is defined by Eq. (4.51) and H
e
/V
+
j
for j = 1,2 are:
H
e
V
+
j
=
F
+
3
V
+
j
F
2
F
3
F
+
2
V
+
j
,
F
+
1
V
+
j
F
3
F
1
F
+
3
V
+
j
,
F
+
2
V
+
j
F
1
F
2
F
+
1
V
+
j
[G]
T
+
+
F
+
3
F
2
F
3
F
+
2
, F
+
1
F
3
F
1
F
+
3
, F
+
2
F
1
F
2
F
+
1
G
+
V
+
j
T
, (4.81)
where
[G]=
(
[G
1
],[G
2
],[G
3
]
)
and G/V =
(
G
1
/V,G
2
/V,G
3
/V
)
.
Equation
(4.80) yields an equivalent expression for the secondary bifurcation.
Proposition 4.4.5. A state V
belongs to the bifurcation manifold for the family i
when there exists a state V
+
such that:
V
+
∈RH(V
) with
ˆ
v
+
(V
;V
+
)=
ˆ
λ
i,+
(V
+
) and
i
(V
+
) ·[G]=0. (4.82)
Proof: We drop the family index i. Assume that
(4.80) is satisfied. Rearranging
the terms, we can rewrite
H
e
/V
+
j
for j = 1,2:
H
e
V
+
j
=
F
3
[G
2
] F
2
[G
3
]
F
+
1
V
+
j
F
3
F
+
2
F
+
3
F
2
G
+
1
V
+
j
+
F
1
[G
3
] F
3
[G
1
]
F
+
2
V
+
j
F
+
3
F
1
F
3
F
+
1
G
+
2
V
+
j
+
F
2
[G
1
] F
1
[G
2
]
F
+
3
V
+
j
F
+
1
F
2
F
1
F
+
2
G
+
3
V
+
j
. (4.83)
Notice that
λ
+
= u
+
ˆ
λ
+
, so the matrix
W
(uF)
λ
W
G , is written at (V
+
,u
+
) as:
u
+
F
+
1
V
+
1
ˆ
λ
+
G
+
1
V
+
1
u
+
F
+
1
V
+
2
ˆ
λ
+
G
+
1
V
+
2
F
+
1
u
+
F
+
2
V
+
1
ˆ
λ
+
G
+
2
V
+
1
u
+
F
+
2
V
+
2
ˆ
λ
+
G
+
2
V
+
2
F
+
2
u
+
F
+
3
V
+
1
ˆ
λ
+
G
+
3
V
+
1
u
+
F
+
3
V
+
2
ˆ
λ
+
G
+
3
V
+
2
F
+
3
(4.84)
78
Setting
ˆ
v
+
(V
;V
+
)=
ˆ
λ
+
(V
+
) in (4.84) and assuming that D
ij
(V
,V
+
)=F
i
[G
j
]
F
j
[G
i
] = 0 for all {i, j}∈P, we use Eqs. ( 4.52) and (4.54) to write an equivalent but
convenient expression for
ˆ
v
+
in each matrix element, obtaining:
u
+
F
+
1
V
+
1
F
3
F
+
2
F
+
3
F
2
F
3
[G
2
]F
2
[G
3
]
G
+
1
V
+
1
u
+
F
+
1
V
+
2
F
3
F
+
2
F
+
3
F
2
F
3
[G
2
]F
2
[G
3
]
G
+
1
V
+
2
F
+
1
u
+
F
+
2
V
+
1
F
+
3
F
1
F
3
F
+
1
F
1
[G
3
]F
3
[G
1
]
G
+
2
V
+
1
u
+
F
+
2
V
+
2
F
+
3
F
1
F
3
F
+
1
F
1
[G
3
]F
3
[G
1
]
G
+
2
V
+
2
F
+
2
u
+
F
+
3
V
+
1
F
+
1
F
2
F
1
F
+
2
F
2
[G
1
]F
1
[G
2
]
G
+
3
V
+
1
u
+
F
+
3
V
+
2
F
+
1
F
2
F
1
F
+
2
F
2
[G
1
]F
1
[G
2
]
G
+
3
V
+
2
F
+
3
(4.85)
Since
H
e
/V
+
j
= 0 for j = 1,2, from Eq (4.83) , it follows that for given by
=
F
3
[G
2
] F
2
[G
3
], F
1
[G
3
] F
3
[G
1
], F
2
[G
1
] F
1
[G
2
]
, (4.86)
the inner products of columns 1 and 2 by
are zero. Since at (V
,V
+
) the expression
H
e
(V
) vanishes, after some calculations, we obtain:
· (F
+
1
, F
+
2
, F
+
3
)=0.
First let us consider the case D
ij
(V
,V
+
) = 0 for all {i, j}∈P, so it is clear that
is a left eigenvector of the system. One can prove that all eigenvectors for
λ
(V
+
)=
v
s
(V
,V
+
) are parallel to . Notice that · ([G
1
],[G
2
],[G
3
]=0, because this equality
satisfies the RH locus
H
e
= 0, with H
e
given by (4.51).
If
D
ij
(V
,V
+
)=0 for a pair {i, j }∈P, using Lemma 4.4.1, the relationships
F
+
i
F
j
F
+
j
F
i
= 0 or D
ij
(V
,V
+
)=0 for all {i, j }∈P are satisfied.
Assume that
D
ij
(V
,V
+
) = 0 for some {i, j}∈P. For concreteness we set i = 3
and j
= 2; the other cases can be proved similarly. Since F
1
[G
3
] F
3
[G
1
] = 0 and
the matrix
(4.85) has the form (a
ij
) for i, j = 1,2,3, and we can write a
11
, a
12
and a
13
,
respectively as:
u
+
F
+
1
V
+
1
F
+
3
F
1
F
3
F
+
1
F
1
[G
3
] F
3
[G
1
]
G
+
1
V
+
1
, u
+
F
+
1
V
+
2
F
+
3
F
1
F
3
F
+
1
F
1
[G
3
] F
3
[G
1
]
G
+
1
V
+
2
and F
+
1
.
Substituting
in (4.86) by the the following vector:
=
0, F
1
[G
3
] F
3
[G
1
], F
2
[G
1
] F
1
[G
2
]
, (4.87)
it is easy to prove that this vector
is in the kernel of the transpose of the matrix
(4.85).
Finally assume that
D
ij
= 0 for all {i, j}∈P. Since is a left eigenvector of the
matrix
(4.85) it follows that ·(F
+
1
, F
+
2
, F
+
3
)=0. From Eq. (4.67.a) in the Proposition
4.4.3, we know that
([G
1
],[G
2
],[G
3
]) = (
ρ
1
F
+
1
,
ρ
1
F
+
2
,
ρ
1
F
+
3
) for any constant
ρ
1
R, so:
· [G]= ·(
ρ
1
F
+
)=
ρ
1
· F
+
= 0.
The converse can be proved similarly by reversing the calculations.
79
Coincidence loci
There are two important types of speed coincidence: coincidence between eigenvalues
and coincidence between eigenvalues and shock speeds. These structures are impor-
tant because the Riemann solution changes when the L (or R) data is prescribed in
different regions relatively to these curves.
Coincidence between eigenvalues in each physical situation. Notice that it is pos-
sible to analyze the system in the variables V without taking into account the Darcy
speed u. At the coincidence locus between eigenvalues the Darcy speeds also coincide,
so the coincidence in the V space consists of the states in a certain physical situation
such that
ˆ
λ
1
(V)=
ˆ
λ
2
(V).
Double contact curves
It consists of the states V
for which there is a state V
+
such that the shock joining
V
and V
+
has speed coinciding with the characteristic speed of family i at () and
of family j at
(+):
V
+
∈RH(V
) with
ˆ
λ
i,
(V
)=
ˆ
v
(V
,V
+
) and
ˆ
v
+
(V
,V
+
)=
ˆ
λ
j,+
(V
+
).
Notice that V
and V
+
can be in the same or in different physical situations.
Inflection curves
The rarefaction curves are useful to construct rarefaction waves where the eigenvalue
varies monotonically; the inflection curve is the curve where the monotonicity fails,
thus rarefaction curves stop at this curve. Since u varies along the rarefaction wave,
it is impossible to disregard the effect of u. Any state W
=(V,u) on the inflection
curve satisfies:
λ
i
(W) ·r
i
(W)=0. (4.88)
However, the Darcy speed can be isolated in Eq. (4.88
), so it is easy to prove that in
the space of primary variables, the inflection curve of family i, for i
= 1,2, consists of
the states V satisfying the equation:
V
ˆ
λ
i
·
ˆ
r
=
ˆ
λ
i
η
3
, (4.89)
where
η
i
(V) for i = 1,2,3 are the coordinates for the eigenvector r given in Eq.
(4.38.b).
Interior boundary contact (extension of the boundary)
Because of the presence of physical region boundaries it is important to obtain the
states V joined to points V
on the physical boundary by shock waves that are charac-
teristic at V for the family i. These states satisfy:
V
∈RH(V
) with V
on the boundary and
ˆ
λ
i
(V)=
ˆ
v
(V,V
).
80
4.4.4 Admissibility of shocks
Frequently, for a fixed state V
there are points in the RH curve of V
that separate
potentially admissible shocks from non-admissible shocks. At these points the shock
speed coincides with the characteristic speed and the shock speed has an extremum,
see Proposition 4.4.4, so the Riemann solution usually changes. There are two type of
such shocks: left characteristic and right characteristic shocks. In this structure, the
shock has speed equal to the characteristic speed at the left or at the right state.
Left characteristic shock.
For a fixed state V
, these points are the V
+
such that:
V
+
∈RH(V
) with
ˆ
v
+
(V
;V
+
)=
ˆ
λ
j,+
(V
+
) for family j = 1 or 2.
Notice that V
and V
+
can be in the same or different physical situations.
Right characteristic shock.
For a fixed V
+
, these points are the V
such that:
V
∈RH(V
+
) with
ˆ
v
(V
,V
+
)=
ˆ
λ
j,+
(V
+
) for family j = 1 or 2.
Notice that V
and V
+
can be in the same or in different physical situations.
4.5 Elementary waves in the single-phase gaseous
situation
4.5.1 Characteristic speed analysis
We simplify the system (4.21)-(4.23) by substituting (4.21) by ( 4.21) divided by
θ
W
and (4.21) by a certain linear combination of (4.21) with (4.22):
t
ϕψ
gw
T
1
+
x
u
ψ
gw
T
1
= 0, (4.90)
t
ϕ
T
1
+
x
uT
1
= 0, (4.91)
t
ϕ
ˆ
H
r
+
ψ
gw
H
gW
+
ψ
gn
H
gN
+
x
u
(
ψ
gw
H
gW
+
ψ
gn
H
gN
)=0, (4.92)
where we have used the equality
ψ
gn
= 1
ψ
gw
.
We differentiate Eqs. (4.90)-(4.92) and we rewrite it in the form
(4.33) with V =
(
ψ
gw
, T) as:
B
t
ψ
gw
T
u
+ A
x
ψ
gw
T
u
= 0, (4.93)
81
where (indicating the differentiation relative to temperature by
):
B
=
ϕ
T
1
ψ
gw
T
2
0
0
T
2
0
(H
gW
H
gN
)
ˆ
C
r
+
ψ
gw
H
gW
+
ψ
gn
H
gN
0
, (4.94)
A
=
uT
1
u
ψ
gw
T
2
ψ
gw
T
1
0 uT
2
T
1
u(H
gW
H
gN
) u(
ψ
gw
H
gW
+
ψ
gn
H
gN
)
ψ
gw
H
gW
+
ψ
gn
H
gN
; (4.95)
in
(4.94) the constant
ˆ
C
r
is the rock heat capacity divided by
ϕ
, see Appendix A.
Using
(4.94) and (4.95), we notice that (u
ϕλ
) is repeated in the first column of
A
λ
B. Factoring this term, the determinant of A
λ
B is (u
ϕλ
) times the deter-
minant of
T
1
ψ
gw
T
2
(u
ϕλ
)
ψ
gw
T
1
0 (u
ϕλ
)T
2
T
1
(H
gW
H
gN
)(u
ϕλ
)(
ψ
gw
H
gW
+
ψ
gn
H
gN
)
ϕλ
ˆ
C
r
ψ
gw
H
gW
+
ψ
gn
H
gN
,
(4.96)
so an eigenvalue and eigenvector for the system
(4.93) are
λ
c
= u/
ϕ
, r
c
=(1,0,0)
T
, (4.97)
which correspond to fluid transport and the wave for this eigenvalue is a contact
discontinuity. The composition
ψ
gw
changes but the speed and the temperature are
constant. There is a single vector for
λ
c
, because the matrix (4.96) with
λ
=
λ
c
has
rank 3, generically. We denote this eigenvalue by
λ
c
to indicate that in the associated
wave only the composition changes.
We find the other characteristic speed
λ
in (4.96).If
ψ
gw
= 0, we add the first line
to the second line multiplied by
ψ
gw
obtaining:
T
1
00
T
1
( u
ϕλ
)T
2
T
1
(H
gW
H
gN
)(u
ϕλ
)(
ψ
gw
H
gW
+
ψ
gn
H
gN
)
ϕλ
ˆ
C
r
ψ
gw
H
gW
+
ψ
gn
H
gN
.
(4.98)
The determinant of
(4.98) is T
2
times
(u
ϕλ
)(
ψ
gw
H
gW
+
ψ
gn
H
gN
)
ϕλ
ˆ
C
r
+(u
ϕλ
)T
1
(
ψ
gw
H
gW
+
ψ
gn
)H
gN
. (4.99)
Setting (4.99) to zero we obtain:
λ
T
=
u
ϕ
1
ˆ
C
r
T
F
, (4.100)
where using
(4.24), F := F(T) is given by:
F(T)=
ψ
gw
ρ
gW
(T)h
gW
(T)+
ψ
gn
ρ
gN
(T)h
gN
(T)+
ˆ
C
r
T. (4.101)
82
As
ρ
gW
,
ρ
gN
, h
gW
, h
gN
and
ˆ
C
r
are positive and
ψ
gw
and
ψ
gn
are non-negative, F(T) is
always positive. The eigenvector associated to
λ
T
is
r
T
=(0,F, u
ˆ
C
r
). (4.102)
This eigenvector is unique, because substituting
λ
T
in (4.98) we see that the resulting
matrix has rank 2. The rarefaction wave has constant composition
ψ
gw
. Notice that
λ
T
<
λ
c
in the physical range, so we can order these waves: the slowest wave is a
thermal wave, with speed
λ
T
; the fastest is a compositional wave, with speed
λ
c
.
Behavior of the gaseous thermal inflection curve
To obtain the rarefaction curves in the spg, we need to study the sign of
λ
T
·r
T
. After
a lengthly calculation, we obtain from
(4.100) and (4.102)
λ
T
·r
T
=
uC
r
T
2
ϕ
Ξ(
ψ
gw
, T), where Ξ =
ψ
gw
ρ
gW
h

gw
+(1
ψ
gw
)
ρ
gN
h

gn
. (4.103)
As u,
ˆ
C
r
, T and
ϕ
are positive, we need to study the sign of Ξ.
The gaseous thermal inflection curve is denoted by
I
T
; for Γ given by (4.26) it is
defined by:
I
T
=
(T,
ψ
gw
) Γ that satisfy
λ
T
·r
T
= 0
. (4.104)
We plot the physical region and
I
T
in Figure (4.2. a), showing the signs of Ξ. In Figure
4.2.b, we plot the horizontal rarefaction lines associated to
λ
c
and the vertical rar-
efaction lines associated to
λ
T
, see (4.102). The waves associated to
λ
c
are contact
discontinuities.
4.5.2 Shocks and contact discontinuities
Using (4.45) in Eqs. (4.90)-(4.92), the RH condition is:
v
s
ϕ
ψ
+
gw
/T
+
ψ
gw
/T
= u
+
ψ
+
gw
/T
+
u
ψ
gw
/T
, (4.105)
v
s
ϕ
1
/T
+
1/T
= u
+
/T
+
u
/T
, (4.106)
v
s
ϕ
H
+
r
+
ψ
+
gw
H
+
gW
+(1
ψ
+
gw
)H
+
gN
(H
r
+
ψ
gw
H
gW
+(1
ψ
gw
)H
gN
)
=
u
+
(
ψ
+
gw
H
+
gW
+(1
ψ
+
gw
)H
+
gN
) u
(
ψ
gw
H
gW
+(1
ψ
gw
)H
gN
),
(4.107)
where H
±
gW
=
θ
W
h
gW
(T
±
)/T
±
and H
±
gN
=
θ
N
h
gN
(T
±
)/T
±
.
In
(4.105)-( 4.107), there is a contact discontinuity with speed (4.97); the Darcy
speed and the temperature are constant on this wave, i.e., u
= u
and T = T
.
The other shock occurs with
ψ
gw
constant; it is a thermal shock with speed v
s
and
Darcy speed u
+
on the right side given by:
v
s
=
u
T
Υ
(T
; T
+
)(T
+
T
)+T
+
and u
+
=
u
ϕ
Υ(T
; T
+
),
83
Figure 4.2: a)-left: The single-phase gaseous physical region Γ , and inflection locus.
b)-right: Rarefaction curves. The horizontal rarefaction curves are associated to
λ
T
;
we indicate with an arrow the direction of increasing speed. The vertical lines are
contact discontinuity curves associated to
λ
c
, in which
ψ
gw
changes, T and u are con-
stant.
where Υ is defined as follows, with
θ
W
,
θ
N
, H
gW
and H
gN
given by Eq. (4.24):
Υ
=
ψ
gw
θ
W
h
+
gW
h
gW
+(1
ψ
gw
)
θ
N
h
+
gN
h
gN
T
ˆ
H
+
r
ˆ
H
r
+
ψ
gw
θ
W
h
+
gW
h
gW
+(1
ψ
gw
)
θ
N
h
+
gN
h
gN
. (4.108)
The RH and rarefaction curves are contained in the horizontal lines in Figure
(4.2.b).
4.6 Contact discontinuities in the single-phase liq-
uid situation
Eq. (4.31) is linear in the spl, so a unique wave is associated to
λ
W
T
, given by (4.31.b).
This wave is a contact discontinuity and there is no genuine rarefaction wave.
Remark 4.6.1. In the spl there exists a cooling discontinuity between a state with
temperature T
and another state with temperature T
+
. For the Riemann data: T
if
x
< 0 and T
+
if x > 0, the solution is T
if x/t <
λ
W
T
and T
+
if x/t >
λ
W
T
.
84
4.7 Elementary waves in the two-phase situation
4.7.1 Characteristic speed analysis
By means of a procedure similar to that in Section 4.4.1, we write the system (4.27)-
(4.29) in the form (4.33) , with V =(s
g
, T), as:
B
t
s
g
T
u
+ A
x
s
g
T
u
= 0, (4.109)
with
B
=
ϕ
(
ρ
gw
ρ
W
)
ϕρ
gw
s
g
0
ϕρ
gn
ϕρ
gn
s
g
0
ϕ
(H
g
H
W
)
ϕ
(
ˆ
C
r
+ C
W
+ s
g
(C
g
C
W
)) 0
, (4.110)
A
=
u
(
ρ
gw
ρ
W
)
f
g
s
g
u
(
ρ
gw
ρ
W
)
f
g
T
+
ρ
gw
f
g
ρ
W
+ f
g
(
ρ
gw
ρ
W
)
u
ρ
gn
f
g
s
g
u
ρ
gn
f
g
T
+
ρ
gn
f
g
f
g
ρ
gn
u(H
g
H
W
)
f
g
s
g
u
C
W
+ f
g
(C
g
C
W
)+(H
g
H
W
)
f
g
T
f
g
(H
g
H
W
)+H
w
,
(4.111)
where C
g
= H
g
/T and the constant C
W
are the gas and water heat capacities given
in Appendix A.
In order to use Eq.
(4.34.b) for system (4.109), we calculate A
λ
B as:
γ
2
γ
1
u
γ
2
f
g
T
+
ρ
gw
γ
4
ρ
W
+ f
g
γ
2
ρ
gn
γ
1
u
ρ
gn
f
g
T
+
ρ
gn
γ
4
f
g
ρ
gn
γ
3
γ
1
C
W
(u
λϕ
)+(C
g
C
W
)
γ
4
ϕλ
ˆ
C
r
+ u
γ
3
f
g
T
f
g
γ
3
+ H
w
, (4.112)
where we have defined
γ
1
=
γ
1
(s, T, u)=u
f
g
s
g
λϕ
and
γ
4
:=
γ
4
(s, T, u)=uf
g
λϕ
s
g
, (4.113)
γ
2
:=
γ
2
(T)=
ρ
gw
ρ
W
, and
γ
3
:=
γ
3
(T)=H
g
H
W
. (4.114)
Setting
γ
1
= 0 in (4.112), its determinant vanishes, and we find the Buckley-Leverett
characteristic speed and characteristic vector:
λ
s
=
u
ϕ
f
g
s
g
, r
s
=(1,0,0)
T
, (4.115)
On this rarefaction curve T and u are constant, only s
g
changes, hence the subscript
s in
λ
s
and r
s
is of a saturation wave. The eigenvector associated to the eigenvalue
(4.115) is unique: substituting
λ
s
in (4.112) leads to a matrix with rank 2, generically.
85
We factor
γ
1
in first column, for
γ
1
= 0. The other eigenvalue of (4.112) is the root
of
det
γ
2
u
f
g
T
γ
2
+
ρ
gw
γ
4
ρ
W
+ f
g
γ
2
ρ
gn
u
ρ
gn
f
g
T
+
ρ
gn
γ
4
f
g
ρ
gn
γ
3
C
W
(u
λϕ
)+
γ
3
γ
4
ϕλ
ˆ
C
r
+ u
f
g
T
γ
3
f
g
γ
3
+ H
W
= 0. (4.116)
We consider first the case
ρ
gn
= 0, the case
ρ
gn
= 0 is considered later. We divide the
second row by
ρ
gn
and we obtain the eigenvalue:
λ
e
=
u
ϕ
f
g
(
γ
3
ρ
gn
/
ρ
gn
γ
3
)
ρ
W
H
W
(
ρ
gw
ρ
gn
/
ρ
gn
γ
2
)
+ C
W
ρ
W
ρ
W
(C
W
+
ˆ
C
r
)+s
g
(
γ
3
ρ
gn
/
ρ
gn
γ
3
)
ρ
W
H
W
(
ρ
gw
ρ
gn
/
ρ
gn
γ
2
)
,
(4.117)
with eigenvector
r
e
=(
1
γ
1
,u
3
)
T
, (4.118)
where
¯
γ
1
=
f
g
s
g
f
g
(
γ
3
ρ
gn
ρ
gn
γ
3
)
ρ
W
H
W
(
ρ
gw
ρ
gn
ρ
gn
γ
2
)
+ C
W
ρ
W
ρ
gn
ρ
W
ρ
gn
(C
W
+
ˆ
C
r
)+s
g
(
γ
3
ρ
gn
ρ
gn
γ
3
)
ρ
W
H
W
(
ρ
gw
ρ
gn
ρ
gn
γ
2
)
, (4.119)
¯
γ
4
=
C
W
( f
g
s
g
)
ρ
W
ρ
gn
+
ˆ
C
r
f
g
ρ
W
ρ
gn
ρ
W
ρ
gn
(C
W
+
ˆ
C
r
)+s
g
(
γ
3
ρ
gn
ρ
gn
γ
3
)
ρ
W
H
W
(
ρ
gw
ρ
gn
ρ
gn
γ
2
)
, (4.120)
1
= ¯
γ
4
ρ
gn
ρ
W
f
g
(
ρ
gw
ρ
gn
ρ
gn
γ
2
)
ρ
gn
ρ
W
+
f
g
T
, (4.121)
3
=
¯
γ
1
¯
γ
4
ρ
gw
ρ
gn
ρ
gn
γ
2
ρ
W
ρ
gn
, (4.122)
and
γ
2
is given in (4.114) .
In Figs.
(4.5.a) and (4.5.b) we can see that in the region where
λ
s
>
λ
e
, along the
rarefaction waves the temperature, gaseous water saturation and u increase; and in
the region where
λ
s
<
λ
e
along the rarefaction waves the temperature and u increase
and the gaseous water saturation decreases. We utilize the subscript e in
λ
e
, r
e
to
indicate that this eigenpair is associated to an evaporation wave.
Coincidence curve
The coincidence curve between
λ
e
and
λ
s
is:
C
s,e
=
(T,s
g
) tp such that
λ
s
(T,s
g
)=
λ
e
(T,s
g
)
. (4.123)
Differentiating
λ
e
in (4.117) with respect to s
g
, equating it to zero and isolating
f
g
/s
g
, we obtain:
86
Lemma 4.7.1. On the coincidence curve C
s,e
, the derivative
λ
e
/s
g
vanishes.
Thus coincidence between eigenvalues occurs where
λ
e
is a stationary. As in Prop.
5.6.4, we can prove the following Proposition:
Proposition 4.7.1. Lines with fixed T intercept the coincidence curve
C
s,e
twice. At
the lowest value of s
g
,
λ
e
has a minimum and at the higher value
λ
e
has a maximum.
Thus, in the tp, when the temperature T varies two branches of the coincidence
curve are generated as sketched in Figure
(4.4.a).
Proof: We verify this proposition geometrically. First, we define
A =

(H
W
H
g
)
ρ
gn
ρ
gn
(H
W
H
g
)
ρ
W
H
W
(
ρ
gw
ρ
gn
ρ
gn
(
ρ
gw
ρ
W
))
.
(4.124)
One can verify that
A > 0 for all T in the tp physical range. We multiply and divide
λ
e
from Eq. (4.117) by A obtaining:
λ
e
=
u
ϕ
f
g
f
e
s
g
s
e
, where f
e
ρ
W
ρ
gn
C
W
A
and s
e
ρ
W
ρ
gn
C
W
+
ˆ
C
r
A
. (4.125)
For the coincidence points with fixed u, it follows from
(4.115):
f
g
s
g
=
f
g
f
e
s
g
s
e
. (4.126)
Eq.
(4.126) defines the tangency points of the secant from the point (s
e
, f
e
) to the
graph of f
g
. One can prove that s
e
> 1 and f
e
> 1. In Figure 4.3, we plot a possible
point
(s
e
, f
e
). At the first point where
λ
e
and
λ
T
coincide the speed is a minimum and
at the second point the speed is maximum.
Behavior of the two-phase evaporation inflection curve
To obtain the evaporation rarefaction curves we need to study the sign of
λ
e
· r
e
.
After a lengthly calculation, we get
λ
e
·r
e
=
u
ϕ
¯
γ
1
¯
γ
4
s
g
s
e
A
A
ρ
gn
ρ
gn
C
W
(
ρ
gw
ρ
gn
ρ
gn
γ
2
)
, (4.127)
where ¯
γ
1
, ¯
γ
4
are defined in Eqs. (4.119) and (4.120). The expression for A is given by
(4.124) and A
is its derivative with respect to temperature.
One can verify that
A
A
ρ
gn
/
ρ
gn
C
W
(
ρ
gw
ρ
gn
ρ
gn
γ
2
) > 0, so the points where
λ
e
·r
e
vanishes satisfy:
λ
e
=
λ
s
or
f
g
s
g
=
C
W
C
W
+
ˆ
C
r
. (4.128)
87
Figure 4.3: The coincidence between
λ
e
and
λ
s
occurs at the tangency point on the
graph of f
g
from (s
e
, f
e
). Since A > 0, s
e
and f
e
are positive, the first coincidence point
is a maximum while the second point is a minimum for
λ
e
.
l
I
II
l
l
l
l
l
D
l
l tr
0
D
l tr
0
andand
and
P
l
D
l tr
0
and
l
l
l
D
l tr
0
and
l
l
D
l
l tr
0
andand
Figure 4.4: a)-left: Coincidence curve C
s,e
. Relative sizes of
λ
s
and
λ
e
in a piece of
the tp state space. The almost horizontal coincidence curve
λ
e
=
λ
s
is not in scale,
because it is very close to the axis S
g
= 0. b)-right: A zoom of the region below the
coincidence curve. In both figures, all curves form the inflection curve, subdividing tp
situation in four parts.
88
We define the curve I
e
as:
I
e
:= {(T, s
g
) tp satisfying Eq. (4.128.b)}.
We denote the inflection curve in the tp of
λ
e
by I
e
; it is defined by:
I
e
=
(T,s
g
) satisfying (4.128.a) and (4.128.b)
(4.129)
The following Lemma follows from Eqs.
(4.128) and (4.123).
Lemma 4.7.2. The coincidence curve
C
s,e
is contained in I
e
. Moreover I
e
= C
s,e
I
e
.
The solid curves in Fig. 4.4 are the
I
e
, where we also plot the sign of
λ
e
·r
e
.In
Figure 4.5, we draw rarefaction curves in the tp.
300 310 320 330 340 350 360 370
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Temperature /K
Gas Saturation
300 310 320 330 340 350 360 370
0.005
0.01
0.015
0.02
0.025
Temperature /K
Gas Saturation
Figure 4.5: a)-left: The rarefaction curves projected in the plane T, s
g
. The thin curves
without arrows are
C
s,e
. The bold curve indicates an invariant curve for the rarefac-
tion field, which a rarefaction curve that reaches the point P in Fig.
(4.4.b). b)-right:
The rarefaction curves in the regions III and IV shown in the Figure
(4.4.b). The
arrows indicate the direction of increasing speed.
For the case
ρ
gn
= 0 but
ρ
gn
= 0, the matrix (4.116) reduces to
γ
2
u
f
g
T
γ
2
+
ρ
gw
γ
4
ρ
W
+ f
g
γ
2
0
ρ
gn
γ
4
0
γ
3
C
W
(u
λϕ
)+
γ
3
γ
4
+
ϕλ
ˆ
C
r
+ u
f
g
T
γ
3
f
g
γ
3
+ H
W
. (4.130)
The eigenvalue and eigenvector are
λ
e
ρ
gn
=0
=
u
ϕ
f
g
s
g
, r
e
ρ
gn
=0
=
¯
γ
4
ρ
gn
f
g
γ
2
+ 1
,1, ¯
γ
4
ρ
gn
γ
2
. (4.131)
The case
ρ
gn
= 0 and
ρ
gn
= 0, with no nitrogen, is studied in [6 and Chapter 5.
89
4.7.2 Shocks in the two-phase situation
As in the spg, we use Eqs. (4.45) and (4.27)-(4.29) to obtain the RH condition:
v
s
ϕ
(s
+
W
s
W
)
ρ
W
+ s
+
g
ρ
+
gw
s
g
ρ
gw
=
u
+
f
+
w
f
w
u
ρ
W
+ u
+
f
+
g
ρ
+
gw
u
f
g
ρ
gw
,
(4.132)
v
s
ϕ
s
+
g
ρ
+
gn
s
g
ρ
gn
= u
+
f
+
g
ρ
+
gn
u
f
g
ρ
gn
, (4.133)
v
s
ϕ
ˆ
H
+
r
ˆ
H
r
+ H
+
W
H
W
+ s
+
g
(H
+
g
H
+
W
) s
g
(H
g
H
W
)
=
u
+
( f
+
w
H
+
W
+ f
+
g
H
+
g
) u
( f
w
H
W
+ f
g
H
g
). (4.134)
The isothermal branch of the RH curve is obtained for constant temperature T,
i.e., T
+
= T
= T. In this case the shock speed is
v
s
=
u
ϕ
f
w
(s
+
w
, T) f
w
(s
w
, T)
s
+
w
s
w
.
On this wave, the Darcy speed is constant. This is the Buckley-Leverett shock; the
RH and rarefaction curves associated to
λ
s
lie on the same straight line.
The non-isothermal wave is obtained after a lengthly calculation. We fix the left
state and the temperature of the right state, with T
+
= T
. Following [10], we can
obtain the shock speed v
s
and Darcy speed u
+
at the right independently of s
+
g
.We
rewrite
(4.132)-( 4.134) as:
v
s
ϕ
s
+
g
u
+
f
+
g

ρ
+
gw
ρ
W
= v
s
ϕ
s
g
ρ
gw
ρ
W
u
f
w
ρ
W
+ f
g
ρ
gw
+ u
+
ρ
W
,
(4.135)
v
s
ϕ
s
+
g
u
+
f
+
g
ρ
+
gn
= v
s
ϕ
s
g
ρ
gn
u
f
g
ρ
gn
, (4.136)
v
s
ϕ
s
+
g
u
+
f
+
g

H
+
g
H
+
W
= v
s
ϕ
ˆ
H
r
ˆ
H
+
r
H
+
W
+ s
W
H
W
+ s
g
H
g
+ u
+
H
+
W
u
( f
w
H
W
+ f
g
H
g
), (4.137)
where we have used s
w
= 1 s
g
and f
w
= 1 f
g
.
As
ρ
W
>
ρ
gw
and H
W
> H
g
in the temperature physical range and we assume that
ρ
gn
= 0 (the term
ρ
gn
is zero only at the water boiling temperature), we add (4.135)
divided by
ρ
+
W
ρ
+
gw
to (4.136) divided by
ρ
+
gn
. Similarly, we add (4.136) divided by
ρ
+
gn
to (4.137) divided by H
+
W
H
+
g
and we obtain:
90
v
s
ϕ
s
g
(
ρ
gw
ρ
W
)
u
f
w
ρ
W
+ f
g
ρ
gw
+ u
+
ρ
W
ρ
gw
ρ
+
W
=
v
s
ϕ
s
g
ρ
gn
u
f
g
ρ
gn
ρ
+
gn
,
(4.138)
v
s
ϕ
ˆ
H
r
ˆ
H
+
r
H
+
W
+ s
W
H
W
+ s
g
H
g
+ u
+
H
+
W
u
( f
w
H
W
+ f
g
H
g
)
H
+
g
H
+
W
=
=
v
s
ϕ
s
g
ρ
gn
u
f
g
ρ
gn
ρ
+
gn
.
(4.139)
From
(4.138) and (4.139), we obtain v
s
and u
+
in terms of s
g
, T
, u
and T
+
as:
v
s
=
u
ϕ
ρ
+
gn
(Π
2
H
+
W
Π
1
ρ
W
) f
g
ρ
gn
(
ρ
+
gw
H
+
W
H
+
g
ρ
W
)
ρ
+
gn
(Π
4
H
+
W
Π
3
ρ
W
) s
g
ρ
gn
(
ρ
+
gw
H
+
W
H
+
g
ρ
W
)
(4.140)
and
u
+
= u
v
s
ϕ
u
s
g
ρ
gn
ρ
+
gn
(
ρ
+
gw
ρ
W
) Π
4
+ Π
2
f
g
ρ
gn
ρ
+
gn
(
ρ
+
gw
ρ
W
)
/
ρ
W
, (4.141)
where Π
1
(s
g
, T
), Π
2
(s
g
, T
), Π
3
(s
g
, T
; T
+
) and Π
4
(s
g
, T
) are given by:
Π
1
= f
w
H
W
+ f
g
H
g
, Π
2
= f
w
ρ
W
+ f
g
ρ
gw
, (4.142)
Π
3
=
ˆ
H
r
ˆ
H
+
r
H
+
W
+ s
W
H
W
+ s
g
H
g
and Π
4
= s
g
(
ρ
gw
ρ
W
). (4.143)
After finding v
s
and u
+
, we substitute them in (4.133) and divide the resulting
equation by u
+
ρ
+
gn
, rearrange the terms, obtaining an implicit equation for s
+
g
:
f
g
(s
+
g
, T
+
) f
s
+
g
s
=
ˆ
v
s
ϕ
, (4.144)
where
f
=
ˆ
u
ρ
gn
/
ρ
+
gn
f
g
and s
=
ρ
gn
/
ρ
+
gn
s
g
,
with
ˆ
v
= v
s
/u
and
ˆ
u
= u
/u
+
.
As v
s
and u
+
do not depend on s
+
g
, the solution of Eq. (4.144) is the intersection
of the graph of f
g
(s
+
g
) with a linear function of s
+
g
. From the shape of the graph of
f
g
, we see that (4.144) can have one, two or three solutions. Each intersection point
belongs to the Rankine-Hugoniot curve
RH(V
). When T
+
varies each intersection
generates a branch of the Rankine-Hugoniot curve.
This is called the non-isothermal Rankine-Hugoniot curve; some examples are
shown in Figure 4.6.
91
300 310 320 330 340 350 360 370
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Temperature /K
Gas Saturation
(320,0.6)
(370,0.8)
(370,0.5)
(370,0.1)
300 310 320 330 340 350 360 370
0.005
0.01
0.015
0.02
0.025
Temperature /K
Gas Saturation
(320,0.07)
(320,0.04)
(360,0.01)
(336,0.04)
Figure 4.6: a)-left: The RH curves projected in the plane T, s
g
for () points marked.
b)- right: The region shown in Figure
(4.4.b). Each RH locus is formed by a non-
isothermal curve and a vertical line which is the isothermal Buckley-Leverett RH
locus. For the point
(336,0.04) the Rankine-Hugoniot locus reduces to the isothermal
curve only.
4.8 Shocks between regions
We consider now shocks that separate physical situations. In Section 4.4,wegave
the RH condition
(4.45) for shocks between regions. The main feature is that the
accumulation and flux terms have different expressions at each shock side.
4.8.1 Shock between the gaseous and the two-phase situations
For the left state we need to specify steam composition
ψ
gw
, the temperature T
and
the Darcy speed u
. For the right state, we specify the temperature T
+
; the steam
composition is a function of temperature (see
(4.20)) and the Darcy speed is obtained
from the other variables. The RH condition
(4.45) is written as:
v
s
ϕ
ρ
W
S
+
w
+
ρ
+
gw
s
+
g
ψ
gw
ρ
gW
= u
+
ρ
W
f
+
w
+
ρ
+
gw
f
+
g
u
ψ
gw
ρ
gW
, (4.145)
v
s
ϕ
ρ
+
gn
s
+
g
ψ
gn
ρ
gN
= u
+
ρ
+
gn
f
+
g
u
ψ
gn
ρ
gN
, (4.146)
v
s
ϕ
ˆ
H
+
r
ˆ
H
r
+ H
+
w
s
+
w
+ H
+
g
s
+
g
ˆ
H
g
= u
+
H
+
w
f
+
w
+ H
+
g
f
+
g
u
ˆ
H
g
, (4.147)
where
ˆ
H
g
=
ψ
gw
H
gW
+
ψ
gn
H
gN
with H
gW
and H
gN
given by Eq. (4.24).
The system
(4.145)-( 4.147) is similar to (4.132)-(4.134). As in that case, we elimi-
92
nate the saturation and we obtain two linear equations in two unknowns v
s
and u
+
:
v
s
ϕ
ψ
gw
ρ
gW
ρ
W
u
ψ
gw
ρ
gW
+ u
+
ρ
W
ρ
+
gw
ρ
W
=
(
v
s
ϕ
u
)
ψ
gn
ρ
gN
ρ
+
gn
, (4.148)
v
s
ϕ
ˆ
H
r
ˆ
H
+
r
H
+
W
+
ˆ
H
g
+ u
+
H
+
W
u
ˆ
H
g
H
+
g
H
+
W
=
(
v
s
ϕ
u
)
ψ
gn
ρ
gN
ρ
+
gn
, (4.149)
We obtain expressions for v
s
and u
+
as functions of the left state (
ψ
gw
, T
,u
) and
T
+
:
v
s
=
u
ϕ
ψ
gn
ρ
gN
ρ
+
gw
H
+
W
H
+
g
ρ
W
ρ
+
gn
ψ
gw
ρ
gW
H
+
W
H
+
g
ρ
W
Ξ
1
ρ
W
Ξ
2
H
+
W
, (4.150)
u
+
=
u
ψ
gw
ρ
gW
ρ
+
gn
ψ
gn
ρ
gN
(
ρ
+
gw
ρ
W
)
v
s
ϕ
Ξ
2
ρ
W
ρ
+
gn
, (4.151)
where Ξ
1
and Ξ
2
are given by:
Ξ
1
=
ˆ
H
r
ˆ
H
+
r
H
+
W
+
ˆ
H
g
ρ
+
gn
ψ
gn
ρ
gN
γ
+
2
, Ξ
2
=
ψ
gw
ρ
gW
ρ
W
ρ
+
gn
ψ
gn
ρ
gN
γ
+
2
.
After finding v
s
and u
+
, we use Eq. (4.146) to obtain an equation similar to (4.144)
for s
+
g
:
v
s
ϕ
=
f
g
(s
+
g
, T
+
) f
§
s
+
g
s
§
, (4.152)
where
s
§
=
ψ
gn
ρ
gN
/
ρ
+
gn
and f
§
=
ˆ
u
ψ
gn
ρ
gN
/
ρ
+
gn
.
Geometrically, Eq.
(4.152) represents tangency points of the secant from the point
(s
§
, f
§
) to the graph f
g
for fixed T
+
with slope v
s
ϕ
.
The Secondary Bifurcation.
The secondary bifurcation consists of the pairs of states for which the Prop. 4.4.4 fails,
see Sec. 4.4.3. The definition of secondary bifurcation is given by Eqs.
(4.80)-( 4.81).
From Eq.
(4.152), we define for T
,
ψ
gw
in the spg and T
+
,s
+
g
in the tp:
F
(T
,
ψ
gw
,u
; T
+
,s
+
g
,u
+
)=u
+
ρ
+
gn
f
g
(s
+
g
, T
+
) v
s
ϕ
s
+
g
ρ
+
gn
ψ
gn
ρ
gN
+ u
ψ
gn
ρ
gN
.
(4.153)
As v
s
and u
+
do not depend on s
+
g
we obtain a simple expression for F/s
+
g
; for
brevity, we rewrite T
+
as T, s
+
g
as s
g
and u
+
as u:
F
s
g
=
v
s
ϕ
u
f
g
s
g
ρ
gn
,
93
so F/s
g
= 0 yields:
v
s
=
u
ϕ
f
g
s
g
. (4.154)
Since both v
s
and u depend on T, the expression obtained for F/T is complicated; a
lengthly calculation yields:
F
T
=
v
s
T
ϕ
s
g
ρ
gn
ψ
gn
ρ
gN
+ v
s
ϕ
s
g
ρ
gn
u
f
g
T
ρ
gn
u
T
f
g
ρ
gn
uf
g
ρ
gn
. (4.155)
Denoting the numerator and denominator of v
s
ϕ
/u
in Eq. (4.150) by n
vs
and d
vs
,
respectively, we obtain:
v
s
T
=
u
ϕ
(
n
vs
/T
)
d
vs
n
vs
(
d
vs
/T
)
(d
vs
)
2
, (4.156)
where d
vs
/T =:
ˆ
H
r
ˆ
H
r
H
W
+
ˆ
H
g
ρ
gn
(C
r
+ C
W
)
ρ
gn
+
ψ
gn
ρ
gN
ρ
gw
(1 H
w
)
ρ
gn
ψ
gw
ρ
gw
ρ
W
Ξ
2
C
W
,
n
vs
T
=
ψ
gn
ρ
gN
ρ
gw
H
W
+
ρ
gw
C
W
C
g
ρ
W
ρ
gn
ψ
gw
ρ
gw
H
W
H
g
ρ
w
ρ
gn
ψ
gw
ρ
gW
C
W
C
g
ρ
W
.
Denoting the numerator of u in Eq.
(4.151) by N
u
, we obtain that:
u
T
=
(
n
u
/T
)
ρ
gn
ρ
W
N
u
ρ
gn
ρ
W
(
ρ
gn
ρ
W
)
2
, (4.157)
where
n
u
= u
ψ
gw
ρ
gW
ρ
gn
ψ
gN
ρ
gw
v
s
T
ϕ
Ξ
2
v
s
ϕ
Ξ
2
T
.
4.8.2 Shock between the two-phase and the liquid situations
The spl can be considered a limit case of the tp situation, with s
w
= 1 and s
g
= 0.
The left state is
(s
g
, T
,u
); in the right state, we specify only the temperature T
+
;
the Darcy speed is obtained from the other variables and the steam composition is a
function of temperature (see
(4.20)). The RH condition (4.45) is written as:
ϕ
v
s
ρ
W
ρ
W
s
W
ρ
gW
s
g
= u
+
ρ
W
u
ρ
W
f
w
u
ρ
gw
f
g
, (4.158)
ϕ
v
s
ρ
gn
s
g
= u
ρ
gn
f
g
, (4.159)
ϕ
v
s
ˆ
H
+
r
ˆ
H
r
+ H
+
w
H
g
s
g
s
w
H
w
= u
+
H
+
w
u
H
w
f
w
+ H
g
f
g
. (4.160)
94
From (4.159), we obtain that
v
s
=
u
ϕ
f
g
s
g
, (4.161)
Substituting v
s
given by (4.161) in Eq. (4.158) we obtain :
u
+
= u
. (4.162)
There is no mass transfer between the tpsituation and the spl situation, so we can
consider the spl as a particular case of the tp.
Lemma 4.8.1. If there is a shock between the tp
and the spl + with T
+
= T
(i.e., a non-isothermal shock), then
f
g
s
g
=
C
W
ˆ
C
r
+ C
W
. (4.163)
Proof: We rewrite
(4.160) using (4.161) and (4.162) as:
H
+
r
+ H
+
W
u
f
g
s
g
u
H
+
W
=
u
f
g
s
g
H
r
+ s
W
H
W
+ s
g
H
g
u
( f
w
H
W
+ f
g
H
g
),
manipulating this equation we obtain
H
+
r
+ H
+
w
u
f
g
s
g
u
H
+
w
=
u
f
g
s
g
H
r
+(1 s
g
)H
w
+ s
g
H
g
u
( f
w
H
w
+ f
g
H
g
),
H
+
r
H
r
+ H
+
w
H
w
f
g
s
g
=
H
+
w
H
w
,
f
g
s
g
=
(
H
+
w
H
w
)
H
+
r
H
r
+ H
+
w
H
w
=
C
W
ˆ
C
r
+ C
W
,
where we have used that H
w
= C
W
(T T
0
) and
ˆ
H
r
=
ˆ
C
r
(T T
0
).
Remark 4.8.1. We call the () states that satisfy (4.163) the condensation curve;
on this curve there is non-isothermal mass transfer between the gaseous and liquid
phases. From Lemma 4.8.1 and Eq. (4.128), we obtain the following:
Corollary 4.8.1. In the tp, the inflection curve
I
e
is the union of coincidence locus C
s,e
with the “condensation curve”.
Lemma 4.8.2. In the set of
() states that satisfy Eq. (4.163), the evaporation char-
acteristic speed in the tp equals the contact discontinuity speed in liquid situation,
i.e.,
λ
e
=
u
b
ϕ
C
W
C
W
+
ˆ
C
R
.
95
Proof: Since f
g
/s
g
satisfies Eq. (4.163), we substitute f
g
= s
g
C
W
/(C
W
+
ˆ
C
r
) in Eq.
(4.117). After some calculations we obtain the result.
Proposition 4.8.1. The Riemann solution for any
() state, ()=(s
, T
,·) that
satisfies Eq. (4.163) and
(+) states, (+) = (0, T
+
,·) in the spl consists of a shock
between
() and (+).
Proof: From Lemma 4.8.1, if the
() belongs to the condensation curve, the shock
speed between regions is
v
s
=
u
ϕ
C
W
ˆ
C
r
+ C
W
,
which is the same as the the contact discontinuity speed in the spl.
The following corollary follows immediately:
Corollary 4.8.2. The Riemann solution for left states L
=(s
L
, T
L
,·) in regions III
or IV (see Fig. 4.4.b) and right states R
=(0,T
R
,·) in the spl consists of a Buckley-
Leverett shock from L to
(0,T
L
,·) followed by a thermal contact discontinuity in the spl
with speed
λ
W
T
given by (4.31) and u
L
= u
R
.
4.9 The Riemann solution for geothermal energy re-
covery
We show an example of Riemann solution. We consider the injection of a two-phase
mixture of water, steam and nitrogen into a rock containing superheated steam (
ψ
R
=
1) at a temperature T
R
> T
b
:
s
g
,
ψ
(T), T,u
L
if x = 0 (the injection point), with u
L
> 0
s
g
= 1,1, T,·
R
if x > 0
(4.164)
From Prop. 4.4.1, the projection of the solution into the primary variables does not
depend on u
L
. We can subdivide the tp region in 4 subregions, relatively to the left
state L.ForL in regions I and II of Figure
(4.4.a) there are 3 subregions with the
following property: for any L in a given subregion the Riemann problem with data R
of form
(4.164) has the same sequence of waves, see Figure 4.7.In4.9.3, we utilize
results of Section 5.6.2 to obtain the Riemann solution.
4.9.1 Subdivision of tp
For the data (4.164) with T
R
fixed, the Riemann solutions have the same wave se-
quence and structure in for L each subregion labelled
L
1
to L
4
of Figure 4.7; subre-
gions III and IV will be divide in Chapter 6. A very important task is to obtain the
curves that bound subregions: the curve E
2
, see Fig. (4.7. a) and Sec. 4.9.1; the curve
E
1
, see Figs. (4.7.a) and (4.7.b) and Sec. 4.9.1; the curve E
3
, see Fig. (4.7.a) and Sec.
4.9.1.
96
L
2
3
L
L
4
L
Figure 4.7: a)-left The 3 subregions in I and II for a Riemann data of form (4.164).
The curves are defined in Secs. 4.9.1, 4.9.1 and 4.9.1. b)-right Zoom of regions II to
IV. Each curve is explained in Section 4.9.1
Remark 4.9.1. The evaporation wave speed
λ
e
for states in L
2
and L
3
in tp is larger
than the thermal wave speed in the spg. From geometrical compatibility, at the right
of the evaporation shock there is no intermediate thermal wave (rarefaction or shock)
in the spg, i.e., the shock between regions should reach a state of the form
(1,
ψ
gw
, T
R
).
Curve E
2
This curve is defined as the states V
in the tp such that the evaporation rarefaction
speed
λ
e
(V
) equals the speed v
BG
(V
;V
+
) of the shock to V
+
=(1,T
R
,
ψ
gw
) in the
spg. Notice that this curve is an extension of the boundary, see Section 4.4.3.
We obtain
ψ
gw
numerically. For each V
state of the form (S ,
ψ
gw
(T), T) in the tp,
we search the state in the RH locus starting at V
with right temperature T
R
. These
restrictions yield a state in the spg with steam composition
ψ
gw
as unknown.
Curve E
1
We are interested in the Riemann solution for left states in I or II. From geometrical
compatibility, to solve the Riemann problem we choose first the slowest wave when-
ever possible. In II, the slowest wave is the evaporation wave. Since we are interested
in connecting a left state at lower temperature to a state at higher temperature, we
use an evaporation rarefaction curve instead of a shock, see Section 4.7. We can do
this until the speed of the evaporation rarefaction curve equals the speed of the shock
between regions; this occurs when the rarefaction curve crosses E
2
.
The integral curves starting at certain left states do not cross E
2
, rather they
reach directly the steam region at boiling temperature. We define the curve E
1
as the
97
evaporation rarefaction curve that crosses E
2
at water boiling temperature. Notice
that for left state in II above E
1
, the evaporation rarefaction curve always crosses E
2
.
Curve E
3
For states V
above the curve E
2
, the evaporation rarefaction speed is larger than
v
BG
(V
;V
+
), for V
+
=(1,T
R
,
ψ
gw
); in Figures (4.4. a) and (4.8. a) we show the rela-
tives sizes of v
BG
,
λ
e
and
λ
s
. When s
g
tends to 1 the Buckley-Leverett shock speed is
zero, so there are transition curves, v
BL
= v
BG
, v
BL
= v
T
and v
BG
= v
T
. From the
Prop. 4.4.2, we obtain a bifurcation curve E
3
where v
T
= v
BL
= v
BG
. For states above
this curve, the Buckley-Leverett shock is slower than v
BG
and v
T
, so there is no direct
shock between the tp and the spg situations.
300 310 320 330 340 350 360 370
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Temperature /K
Gas Saturation
E
3
E
2
E
1
λ
e
=λ
s
λ
e
=λ
s
λ
s
=v
BG
λ
s
<v
BG
λ
s
>v
BG
300 310 320 330 340 350 360 370
0
0.005
0.01
0.015
0.02
0.025
0.03
Temperature /K
Gas Saturation
II
III
IV
λ
e
=λ
s
B
Figure 4.8: a)-left) The relevant curves for the Riemann solution for the left states in
the tp situation. b)-right) The curve B is a bifurcation curve; for left states V out of
this curve, the shock and rarefactions starting in V do not cross B.
4.9.2 The Riemann solution
We use the notation V
L
=
s
g
,
ψ
gw
(T), T
L
and V
R
=
s
g
= 1,1,T
R
. All the variables
for the intermediate states are written out. The states are labelled by 1, 2, etc. We
use the following nomenclature: evaporation rarefaction is R
E
, the Buckley-Leverett
shock and rarefaction are S
BL
and R
BL
, the shock between the tp and the spg is S
BG
and the contact compositional discontinuity is S
C
. We use the notation for wave se-
quences established in Section 2.1.3.
V
L
in L
4
.
There is an evaporation rarefaction curve from V
L
up to V
1
=(s
L
,1,T
b
), where T
b
is
the boiling temperature. For this state there is no nitrogen and so
ψ
gw
= 1. The so-
98
lution after the intermediate state (s
M
,1,T
b
) was found in Chapter 5: an evaporation
shock between the boiling situation to the steam situation appears, denoted by S
VS
.
The solution consists of the waves R
e
R
BL
S
VS
.
V
L
R
e
V
1
R
BL
V
2
S
VS
V
R
. (4.165)
V
L
in L
3
.
There is a rarefaction from V
L
up to V
1
=(s
,
ψ
gw
(T
), T
) in tp; V
1
is a state where
λ
e
(V
1
)=v
BL
(V
1
;V
2
), with V
2
=(1,
ψ
gw
, T
R
) in spg. We have described how to ob-
tain
ψ
gw
in Section 4.9.1. Finally, there is a compositional contact discontinuity at
temperature T
R
with speed v
C
from V
2
to V
R
.
The solution consists of the waves R
E
S
BG
S
C
with sequence:
V
L
R
e
V
1
S
BG
V
2
S
C
V
R
. (4.166)
L
R
R
e
L
1
2
R
R
e
S
BG
S
C
Figure 4.9: Riemann solutions in phase space, omitting the surface shown in Fig.
4.1. a) Left: Solution
(4.165) for V
L
∈L
4
, Sec. 4.9.2. b) Right: Solution (4.166) for
V
L
∈L
3
, Sec. 4.9.2. The numbers 1 and 2 indicate the intermediate states V in wave
sequence.
V
L
in L
2
.
Let V
1
=(1,
ψ
gw
, T
R
) be a state in the spgsituation. Since
λ
e
(V
L
) > v
BL
(V
L
;V
1
), there
is a shock between V
L
and V
1
with speed v
BG
. From the state V
1
there is compositional
contact discontinuity with speed v
C
to V
R
.
The solution consists of the wave S
BG
S
C
with sequence:
V
L
S
BG
V
1
S
C
V
R
. (4.167)
99
V
L
in L
3
.
Since v
BG
and v
BL
are smaller than v
T
, there is a Buckley-Leverett saturation shock
between V
L
and V
1
=(1,
ψ
gw
(T
L
), T
L
). From V
1
there is a thermal shock with speed
v
T
to V
2
=(1,
ψ
gw
(T
L
), T
R
) and finally there is a compositional contact discontinuity
with speed v
C
to V
R
.
The solution consists of the waves S
BL
S
T
S
C
with sequence:
V
L
S
BL
V
1
S
T
V
2
S
C
V
R
. (4.168)
L
1
R
S
BG
S
C
L
2
R
1
S
BL
S
T
S
C
Figure 4.10: Riemann solutions in phase space, removing the surface shown in Fig.
4.1. a) Left: Solution
(4.167) for V
L
∈L
2
, Sec. 4.9.2. b) Right: Solution (4.168) for
V
L
∈L
1
, Sec. 4.9.2. The numbers 1 and 2 indicate the intermediate states V in wave
sequence.
4.9.3 V
L
in III and IV
The integral curves or RH locus starting at any point in III or IV do not reach the wa-
ter boiling temperature. Moreover, the RH locus between the tpand the spgsituations
for points in III or IV do not reach temperatures above the temperature T
L
. So the
only possibility to get to the right state is by crossing the curve B. However, the evap-
oration wave does not cross the curve B; to cross it we need a Buckley-Leverett shock.
The Buckley-Leverett shock for an initial state below B and a final state above B is
faster than the evaporation wave
λ
e
. Moreover, it is faster than the shock between
regions. So with a Buckley-Leverett shock of this type, it is only possible to reach
states of form
(1,
ψ
gw
(T), T) with T < T
b
. Since the thermal wave is slower than the
Buckley-Leverett shock, so it is impossible to construct the Riemann solution satisfy-
ing the geometrical compatibility principle for waves only in the tp situation.
To obtain the Riemann solution, we utilize the results of Section 5.6.2,soitis
postponed to Chapter 6
100
CHAPTER 5
Riemann solution for steam and water flow
In this section, we use part of the theory developed for conservation laws to solve
a system of balance equations for steam and water flow in a porous medium, see
[20], [45], [59] and [60] for hyperbolic systems. The solution exhibits an intriguing
yet systematic structure. It is desirable to obtain a theory for balance equations as
complete as that for conservation laws; combustion phenomena are also modelled by
balance equations, see
[1 ] and references theirein.
This class of balance equations has appeared in mathematical models for clean-up,
see [6]. Soil and groundwater contamination due to spills of non-aqueous phase liq-
uids (NAPLs) have received a great deal of attention from society, because, in general,
these components can cause damage to the ecosystem and environmental impact to
a large area around the spills. Removal of contaminants with steam is considered
to be an attractive groundwater remediation technique. We consider here a model
for steam injection presented in
[6 ]. Steam injection is widely studied in Petroleum
Engineering, see [6] and references theirein.
In the recovery of geothermal energy, a steam producing well crosses porous rock
containing water at high pressure and temperature. The water usually boils in the
rock, forming a steam vaporization front.Because of the high pressures and pressure
gradients, the nature of the flow equations changes: for single phase steam or water
flow the asymptotic solution depends on x
/
t, see [37]. In many situations, a two-
phase region develops, precisely as in the case analyzed in our work for small pressure
gradients. The precise asymptotic behavior for for flow of water and steam containing
a two-phase region seems to be an open problem flow under high pressures. Tsypkin
et. al.,
[66]. Studies in steam and water injection is a related application in several
works: Bruining
[6 ], K ondrashov [ 35], Kulikoskii [37], [38] and Tsypkin et. al. [63]-
[67].
We consider the constant rate injection of a mixture of steam/water in a specified
proportion into a porous medium filled with another homogeneous steam/water mix-
101
ture. We study all possible proportions of steam and water as boundary and initial
conditions for the problem. We present a physical model for steam injection based
on mass balance and energy conservation equations. We present the main physical
definitions and equations; we refer to
[6 ] for more details. We study the three possible
physical phase mixture situations: single-phase gaseous situation, which represents a
region with superheated steam, called steam region, sr; a two-phase situation, which
represents a region where the water and steam coexist called boiling region, br; and a
single-phase liquid situation which represents a region with water only, called water
region, wr. We reduce the three balance equations system presented in Sec. 5.1 to a
system of conservation laws of form
(2.2):
t
G
(V)+
x
uF
(V)=0,
where V
=(V
1
) : R × R
+
−→ R represents the variables to be determined;
G
=(G
1
, G
2
) : −→ R
2
and F =(F
1
, F
2
) : −→ R
2
are the accumulation vector
and the flux vector, respectively; u :
R ×R
+
−→ R , u = u(x, t) is the total velocity.
It is useful to define U
=(V,u). The vector V represents the water saturation s
w
and
the temperature T. The state of the system is represented by
(s
w
, T,u). Eq. (2.2)
has an important feature, the variable u does not appear in the accumulation term, it
appears isolated in the flux therm, therefore this equation has an infinite speed mode
associated to u. Nevertheless we are able to solve the complete Riemann problem
associated to Eq.
(2.2). Moreover, under certain hypotheses it is possible to solve
numerically the problem, in
[40] Lambert et. al. consider a model in the balance form
for nitrogen and steam injection.
In
[6 ], Bruining et. al. considered as initial condition for the Riemann problem,
a porous rock filled with water at a temperature T
0
, in which a mixture of water
and steam at saturation temperature (boiling temperature) in given proportions is
injected. The main feature was the existence of a Steam Condensation Front (SCF),
which is a shock between the br and the wr. The analysis of the shock between each
pair of regions is important because bifurcations occur and frequently non-classical
structures appear in the solution.
In this work, we completely solve the Riemann problem. We study the three possi-
ble physical phase mixture situations: single-phase superheated steam gaseous situa-
tion, in the sr; a two-phase situation where the water and steam coexist in equilibrium
in the br; and a single-phase wr.
The Riemann problem A is the injection of a mixture of water and steam at boiling
temperature in a porous rock filled with steam at temperature above the boiling tem-
perature (superheated steam). In this case, a new wave, a vaporization shock (VS),
appears between the br and the sr. In the Riemann problem B, we inject liquid water
in a porous rock containing water and steam at boiling temperature. These initial
and boundary conditions are the reverse of those considered in
[6 ]. There is a water
evaporation shock (WES) between the sr and the br. In the Riemann problem C,we
inject superheated steam in a porous rock containing water and steam at boiling tem-
perature. There appears a condensation shock (CS) between the sr and the br. This
102
Riemann solution is the most interesting solution; it has a rich bifurcation structure.
We obtain two bifurcation curves. The first bifurcation is the TCS locus, where the
left thermal characteristic speed in the sr coincides with the condensation shock speed
v
CS
. The second bifurcation is the CSS locus, where the condensation shock speed v
CS
coincides with the right saturation characteristic speed
λ
b
s
in the br. These two bi-
furcation curves intersect at a point, the double bifurcation SHB (see Fig. 5.8). This
state is very important because it is an organizing point between several different
phase mixtures. In the Riemann problem D, we inject water at a temperature below
its boiling temperature in a porous rock containing superheated steam. There is a br
between the wr and the sr. So the Riemann solution consists of a combination of the
waves in the Riemann problem B and C. In the Riemann problem E, we inject super-
heated steam in a porous rock containing water below its boiling temperature. As in
the Riemann problem D, there is a br between the wr and the sr. For this Riemann
problem, the solution is obtained combining the Riemann solution B and the solution
F , see
[6 ] and [39].
In Sec. 5.1, we present the mathematical and physical formulations of the injection
problem in terms of balance equations. In Sec. 5.2, we consider separately each region
in different physical situations under thermodynamical equilibrium and we rewrite
the corresponding balance equations in conservative form. In Sec. 5.4, we study the
shock and rarefaction waves that occur in each physical situation separately. In Sec.
5.5, we study the shocks in the transitions between regions. In Sec. 5.6, we present
the solution of the Riemann problem for the five types of injection. In Appendix A, we
describe notation and physical quantities appearing in the physical model.
5.1 Mathematical and Physical model
5.1.1 The model equations
Ignoring diffusive effects, the mass balance equation for liquid water and steam read:
t
ϕρ
w
s
w
+
x
ρ
w
u
w
=+q, (5.1)
t
ϕρ
g
s
g
+
x
ρ
g
u
g
= q, (5.2)
where
ϕ
is the rock porosity assumed to be constant; s
w
and s
g
are the water and
steam saturations;
ρ
w
is the water density, which is assumed to be constant for sim-
plicity; the steam density
ρ
g
is a function of the temperature T (i.e, we neglect the
effects of gas compressibility) and decreases with temperature; the term q is the mass
transfer between the gaseous and liquid water; u
w
and u
g
are the water and steam
phase velocity.
Disregarding heat conductivity, the energy balance equation can be written as:
t
ϕ
(
ˆ
H
r
+
ρ
w
h
w
s
w
+
ρ
g
h
g
s
g
)+
x
(u
w
ρ
w
h
w
+ u
g
ρ
g
h
g
)=0, (5.3)
103
where H
r
is the rock enthalpy per unit volume and h
w
and h
g
are the water and gas
enthalpies per unit mass, respectively, and
ˆ
H
r
= H
r
/
ϕ
. The enthalpies and
ρ
g
are
functions only of temperature and their expressions are found in Appendix A. From
these expressions, one can see that the enthalpies are increasing functions and that
h
g
is a convex function.
Remark 5.1.1. In
[6], Bruining et. al have defined the water and gas enthalpies per
unit volume as:
H
w
=
ρ
w
h
w
and H
g
=
ρ
g
ρ
g
. (5.4)
From the expressions for
ρ
w
,
ρ
g
, h
w
and h
g
in Appendix A, we can prove that there is a
temperature
T
< T
b
where H
w
and H
g
satisfy:
(i) H
w
(
T
)=H
g
(
T
);
(ii ) H
w
(T) < H
g
(T) if T <
T;
(iii ) H
w
(T) > H
g
(T) if T >
T.
The gas and water enthalpies per unit volume increase when the temperature in-
creases. Moreover, h
g
> h
w
for all T in the physical range.
5.1.2 Physical Model
To determine the fluid flow rate, we use Darcy’s law for multiphase flow without
gravity and capillary pressure effects:
u
w
=
kk
rw
µ
w
p
x
, u
g
=
kk
rg
µ
g
p
x
, (5.5)
where k is the absolute permeability for rock (see Appendix A); the relative perme-
ability functions k
rw
and k
rg
are considered to be power functions of their respective
effective saturations (see Appendix A);
µ
w
and
µ
g
are the viscosity of liquid water and
the viscosity of steam and they are functions of temperature; p is the common pres-
sure of the liquid and gaseous phase. We define the fractional flow functions for water
and steam depending on the saturation and temperature as follows, see Figure 5.1:
f
w
=
k
rw
/
µ
w
d
, f
g
=
k
rg
/
µ
g
d
where d
=
k
rw
µ
w
+
k
rg
µ
g
. (5.6)
Using
(5.6) in Darcy’s law (5.5)
u
w
= uf
w
, u
g
= uf
g
, where u = u
w
+ u
g
, (5.7)
and u is the total or Darcy velocity. The saturations s
w
and s
g
add to 1.
104
Increasing T
00
10
-4
T/K500300
F
/(Pa.s)
w F
2x10
-5
T/K500300
/(Pa.s)
g
Figure 5.1: a) Left: The shape of water fractional flux f
w
, originating from typical
k
rw
(s
w
) and k
rg
(s
g
), given in (B.6) for different values of the temperature T. The
separation between the curves is tiny in reality. b) Center: Water viscosity
µ
w
(T). c)
Right: Gas viscosity
µ
g
(T).
5.2 Regions under thermodynamical equilibrium
As we will see later, the five zones can be organized in three regions in different
physical situations where the fluids are in thermodynamic equilibrium, which is of-
ten specified by an equation of state (EOS). Each physical situation determines the
structure of the governing system of equations. One region consists of steam only,
with temperature at least T
b
(the condensation temperature of pure water, which is
around 373.15K at atmospheric temperature), where we must determine two vari-
ables: temperature and Darcy velocity u. The steam saturation is s
g
= 1 (s
w
= 0).
There is a second region consisting of steam and water, with liquid water saturation
s
w
and gas saturation s
g
both less than 1. We must determine two variables: the
velocity and saturation (either s
w
or s
g
, because they add to 1); the temperature here
is known and its value is T
= T
b
. Finally there is a region of liquid water, where
we must also determine two variables: temperature and velocity. The saturation is
known: s
g
= 0 and s
w
= 1.
We summarize these regions as follows:
s
w
\T T > T
b
T = T
b
T < T
b
s
w
= 0 superheated steam zone hot steam zone 3
s
w
< 1 1 hot steam-water zone 4
s
w
= 1 2 hot water zone cold water zone
Table 1: Classification according to saturation and temperature.
We call “steam region” (represented by sr”) the superheated steam zone. We call
“boiling region” (represented by br) the hot steam zone together with the hot steam-
water zone and the hot water zone. We call “water region” (represented by wr”) the
hot water zone together with the cold water zone. In
[6 ], there is no region with steam
105
above the boiling temperature; the sr and wr are called “hot region” and “liquid water
region” respectively.
Notice that the hot steam zone at boiling temperature belongs to both sr and br;
also, the hot water zone belongs to both br and wr.
Remark 5.2.1. Because of thermodynamical equilibrium, steam cannot exist at tem-
peratures lower than T
b
; similarly, there is no liquid water at a temperature above T
b
.
Thus the regions with numbers 1 -4 in Table 1 do not exist because of our requirement
of thermodynamic equilibrium. Regions 1- 4 would represent the following unstable
mixtures: (1) superheated steam with water, (2) superheated water, (3) steam below T
b
and (4) steam-water below T
b
.
5.3 Equations in conservative form
From the previous discussion, we notice that in each region under thermodynamic
equilibrium there are two variables to be determined in the system
(5.1)-( 5.3); the
other variables are trivial. For example, in the br the temperature and Darcy speed
are determined by the system of equations, but the saturation is trivial, its value is
s
w
= 0. Thus we can rewrite the system (5.1)-(5.3) as a system of two conservation
laws and two variables as follows. We add Eq.
(5.1) to (5.2) and use (5.7):
t
ϕ
ρ
w
s
w
+
ρ
g
s
g
+
x
u
ρ
w
f
w
+
ρ
g
f
g
= 0. (5.8)
Using
(5.7) in the energy conservation equation (5.3), it becomes:
t
ϕ
(
ˆ
H
r
+
ρ
w
h
w
s
w
+
ρ
g
h
g
s
g
)+
x
u
(
ρ
w
h
w
f
w
+
ρ
g
h
g
f
g
)=0. (5.9)
We will use
(5.8)-( 5.9) from now on. Not only this system models the flow in each
region under thermodynamic equilibrium, but it also determines the shocks between
regions (see Sec. 5), when supplemented by appropriate thermodynamic equations of
state.
As initial conditions, we assume that the porous rock is full of a mixture of water
and steam (saturation s
w
(x, t = 0)=s
R
) with constant temperature T(x, t = 0)=T
R
.
As boundary conditions at the injection point at the left of the porous rock, the total
injection rate u
L
is specified as a constant. The constant water-steam injection ratio
needs to be given too, which is
(s
L
, T
L
). It is specified in terms of the water fractional
flow f
w
(s
L
, T
L
) at the injection point.
5.4 Elementary waves under thermodynamical equi-
librium
We consider rarefaction and shocks waves for (5.8)-(5.9) in each region.
106
5.4.1 Steam region -sr
The temperature is high, T > T
b
, so there is only steam, s
w
= 0, (notice that from
Eq.
(5.1), q 0) and the state (s
w
, T,u) can be represented by (0,T,u ). The system
(5.8)-( 5.9) reduces to
t
ϕρ
g
+
x
u
ρ
g
= 0, (5.10)
t
ϕ
(
ˆ
H
r
+
ρ
g
h
g
)+
x
u
ρ
g
h
g
= 0. (5.11)
Rarefaction wave
Assuming that all dependent variables are smooth, we can differentiate
(5.10) and
(5.11) with respect to their variables:
ϕρ
g
T
t
+ u
ρ
g
T
x
+
ρ
g
x
u
= 0, (5.12)
ϕ
ˆ
H
r
+ C
g
T
t
+ uC
g
T
x
+
ρ
g
h
g
x
u
= 0, (5.13)
where prime denotes derivative relative to T.
We use the notation
ˆ
C
r
=
ˆ
H
r
= d
ˆ
H
r
/dT for the effective rock heat capacity di-
vided by
ϕ
and C
g
= (
ρ
g
h
g
)/T for the steam heat capacity per unit volume; we as-
sume that the effective rock heat capacity
ˆ
C
r
is constant (see Appendix A). We rewrite
(5.12)-( 5.13) as:
B
t
T
u
+ A
x
T
u
= 0, (5.14)
where
B
=
ϕρ
g
0
ϕ
(
ˆ
C
r
+ C
g
) 0
and A
=
u
ρ
g
ρ
g
uC
g
ρ
g
h
g
. (5.15)
The characteristic speed
λ
and the eigenvector
r =(r
1
,r
2
)
T
=(dT, du)
T
in the fol-
lowing system are the speed of rarefaction waves and the characteristic direction,
respectively:
det
(A
λ
B)=0, A
r =
λ
B
r. (5.16)
We find only one characteristic speed and vector:
λ
g
T
(T,u)=
u
ϕ
ρ
g
C
g
ρ
g
ρ
g
h
g
ρ
g
(
ˆ
C
r
+ C
g
)
ρ
g
ρ
g
h
g
=
u
ϕ
ρ
g
c
g
ˆ
C
r
+
ρ
g
c
g
, and
r
T
=
1,
u
ˆ
C
r
T(
ˆ
C
r
+
ρ
g
c
g
)
T
,
(5.17)
where
ρ
g
=
ρ
g
(T), h
g
= h
g
(T), C
g
= C
g
(T), and the derivatives relative to temperature
are
ρ
g
=
ρ
g
(T), c
g
= h
g
(T) and
ˆ
C
r
is constant; we used the equality
ρ
g
=
ρ
g
/T which
follows from Eq.
(B.4). The notation for this wave has subscript T because it is a
thermal wave; the saturation (s
w
= 0) stays constant, but the temperature T and the
107
speed u change. We obtain the thermal rarefaction curve in (T,u) space from
r
T
in
(5.17):
du
dT
= u
ˆ
C
r
T(
ˆ
C
r
+
ρ
g
(T)c
g
(T))
or
du
u
=
dT
T
1 +
ρ
g
(T)c
g
(T)/
ˆ
C
r
. (5.18)
Remark 5.4.1. Normally
ρ
g
c
g
ˆ
C
r
, so we can approximate
λ
g
T
by:
λ
g
T
u
ϕ
ρ
g
c
g
ˆ
C
r
. (5.19)
In Appendix A, we obtain a better approximation for the characteristic speed
λ
g
T
.
The rarefaction wave in the
{x, t} plane is the solution of the following equations:
dT
d
ξ
,
du
d
ξ
T
=
r
T
, with
ξ
=
x
t
=
λ
g
T
(u(
ξ
), T(
ξ
)). (5.20)
From Eqs.
(5.17.c) and (5.20.a), dT/d
ξ
satisfies:
dT
d
ξ
= 1,
so we write T as:
T
=
ξ
ξ
+ T
, (5.21)
where, from Eqs
(5.20.b),
ξ
satisfies:
ξ
=
λ
g
T
(u
, T
),
Remark 5.4.2. If we make the same approximation used in Eq. (5.19) from Remark
5.4.1 we obtain for Eq.
(5.18):
du
u
=
dT
T
or
u
T
=
u
T
, (5.22)
where u
and T
are any points in the br, so we obtain that u/T is constant on rar-
efaction wave, i.e., Eq.
(5.22) defines a Riemann invariant. So from (5.21), the speed
u in the plane
{x, t} is given by:
u =
u
T
ξ
ξ
+ T
.
In Appendix A, we obtain a better approximation for the rarefaction wave
(5.18).
Remark 5.4.3. In the sr, the temperature decreases from left to right along the thermal
rarefaction wave. In the Section 5.4.1 we consider a thermal steam shocks; analogously,
on the right of such a shock the temperature is higher than on the left.
108
Thermal steam shock
We assume now that T
+
> T
T
b
. Let us consider the thermal discontinuity with
speed v
g
T
between the () state (0,T
,u
) and the (+) state (0,T
+
,u
+
). For such a
thermal steam shock, Eqs. (5.10)-(5.11) yield the following Rankine-Hugoniot (RH)
condition:
v
g
T
=
u
+
ρ
+
g
u
ρ
g
ϕ
(
ρ
+
g
ρ
g
)
=
u
+
ρ
+
g
h
+
g
u
ρ
g
h
g
ϕ
(
ˆ
H
+
r
+
ρ
+
g
h
+
g
) (
ˆ
H
r
+
ρ
g
h
g
)
, (5.23)
where h
±
g
= h
g
(T
±
),
ˆ
H
±
r
=
ˆ
H
r
(T
±
) and
ρ
±
g
=
ρ
g
(T
±
). From the second equality in Eq.
(5.23), we obtain u
+
as a function of u
:
u
+
= u
(
ˆ
H
+
r
ˆ
H
r
)/
ρ
+
g
+ h
+
g
h
g
(
ˆ
H
+
r
ˆ
H
r
)/
ρ
g
+ h
+
g
h
g
; (5.24)
it is easy to see that the denominator of
(5.24) is positive. Moreover u
+
> u
.
We substitute
(5.24) in Eq. (5.23); since u
+
is function of u
, we obtain v
g
T
=
v
g
T
(T
,u
; T
+
) or v
g
T
= v
g
T
(T
; T
+
,u
+
):
v
g
T
=
u
ϕ
h
+
g
h
g
(
ˆ
H
+
r
ˆ
H
r
)/
ρ
g
+ h
+
g
h
g
=
u
+
ϕ
h
+
g
h
g
(
ˆ
H
+
r
ˆ
H
r
)/
ρ
+
g
+ h
+
g
h
g
. (5.25)
Remark 5.4.4. Notice that we can rewrite
(5.25.a) as:
v
g
T
= v
g
T
(T
,u
; T
+
)=
u
ϕ
h
+
g
h
g
/
(
T
+
T
)
ˆ
C
r
/
ρ
g
+
h
+
g
h
g
/
(
T
+
T
)
.
Defining
λ
±
g
=
λ
g
T
(T
±
,u
±
), from convexity of h
g
(T) it follows that v
g
T
<
λ
g
for T
+
> T
.
Using
(5.25.b) we see that
λ
+
g
< v
g
T
<
λ
g
if T
+
> T
, so the steam shock satisfies the
Lax condition. The equality v
g
T
=
λ
g
holds if, only if, T
+
= T
.
5.4.2 Boiling region - br
Because the temperature is constant and equal to the boiling temperature T
b
it can be
shown (see [6]), that there is no mass exchange between phases and that the system
(5.8)-( 5.9) reduces to a single scalar equation with fixed u = u
b
:
ϕ
t
s
w
+ u
b
x
f
w
= 0. (5.26)
Eq. ( 5.26
) supports classical Buckley-Leverett rarefaction and shock waves.
109
Saturation shocks
Consider a
() state (s
w
, T
b
,u
b
) and a (+) state (s
+
w
, T
b
,u
b
); we obtain the following
RH condition, where u
b
is the common Darcy velocity, T = T
b
and we use the nomen-
clature f
b
w
(s
w
)=f
w
(s
w
, T
b
):
v
b
s
(s
w
,u
b
;s
+
w
)=v
s
(s
w
, T
b
,u
b
;s
+
w
)=
u
b
ϕ
f
b
w
(s
+
w
) f
b
w
(s
w
)
s
+
w
s
w
=
u
b
ϕ
f
b
g
(s
+
g
) f
b
g
(s
g
)
s
+
g
s
g
. (5.27)
A particular shock for (5.26) separates a mixture of steam and water on the left from
pure water on the right, both at boiling temperature. Following
[6 ] we call it the Hot
Isothermal Steam-Water shock (or HISW) between the
() state (s
w
, T
b
,u
b
) and the
(+) state (s
+
w
= 1,T
b
,u
b
). It has speed v
b
g,w
given by:
v
b
g,w
(s
w
,u
b
)=v
g,w
(s
w
, T
b
,u
b
)=
u
b
ϕ
1 f
b
w
(s
w
)
1 s
w
=
u
b
ϕ
f
b
g
(s
g
)
s
g
. (5.28)
Another particular shock for (5.26) separates pure water on the left from a mixture
of steam and water on the right, both at boiling temperature. We call it the Hot
Isothermal Water-Steam shock (or HIWS) between the
() state (s
w
, T
b
,u
b
) and the
(+) state (s
+
w
= 0,T
b
,u
b
). It has speed v
b
g,s
given by
v
b
g,s
(s
w
,u
b
)=v
g,s
(s
w
, T
b
,u
b
)=
u
b
ϕ
f
b
w
(s
w
)
s
w
. (5.29)
Notice that v
g,s
(s
w
= 0,T
b
,u
b
)=v
g,w
(s
w
= 1,T
b
,u
b
)=0.
Saturation rarefaction waves
We will denote by
λ
b
s
the speed of propagation of saturation waves in the br.Itis
obtained from Eq. (5.26) as:
λ
b
s
=
λ
s
(s
w
, T
b
,u
b
)=
u
b
ϕ
f
b
w
s
w
(s
w
). (5.30)
5.4.3 Water region - wr
The system (5.8)-(5.9) reduces to a scalar equation, with constant u
w
and s
w
= 1:
ϕ
t
ˆ
H
r
(T)+
ρ
w
h
w
(T)
+ u
w
x
ρ
w
h
w
(T)=0. (5.31)
Between a
() state (1,T
0
,u
w
) and a (+) state (1,T, u
w
), the following RH condi-
tion for the thermal discontinuity is valid:
v
w
T
=
u
w
ϕ
ρ
w
h
w
h
0
w
ˆ
H
r
+
ρ
w
h
w
(
ˆ
H
0
r
+
ρ
w
h
0
w
)
=
u
w
ϕ
C
w
ˆ
C
r
+ C
w
, (5.32)
110
where u
w
is Darcy speed in the wr and C
w
=
ρ
w
h
w
/T; the second equality is ob-
tained taking into account
(B.5).IfT
0
= T
b
or T = T
b
, then u
b
= u
w
. From (5.32),
the discontinuity is a contact wave and there is no other characteristic speed in this
region.
5.5 Shocks between regions
Within shocks separating regions there is no thermodynamic equilibrium, so q is not
zero; however we can still use the system
(5.8)-( 5.9), because in each region the num-
ber of variable to be determined in the system
(5.1)-( 5.3) is at most 2. This system
contains another variable, namely the mass transfer term q. However this variable
is not essential to obtain the Riemann solution. It is useful to define the cumulative
mass transfer function:
Q(x, t)=
x
q(
ξ
,t)d
ξ
, (5.33)
where this integral should be understood in the distribution sense. From Eq.
(5.33),
we can write q
= Q(x, t)/x.
We also define
Q
(t)=Q(x
,t) and Q
+
(t)=Q(x
+
,t), where x
and x
+
are the
points immediately on the left and right of the transition between regions. We define
the accumulative balance as the difference between
Q
+
(t) and Q
(t) and denote it
by
[Q].
We can rewrite the system
(5.1)-( 5.3) (in distribution sense) as:
t
G
(s
w
, T)+
x
(
uF(s
w
, T) Q(s
w
, T)
)
= 0, (5.34)
where Q
=
(
Q(V), −Q(V),0
)
T
; G =(G
1
, G
2
, G
3
)
T
and F =(F
1
, F
2
, F
3
)
T
. The compo-
nents of F and G are readily obtained from Eqs.
(5.1)-( 5.3) using (5.7).
The shock waves are discontinuous solutions of Eq.
(5.34) and satisfy the RH
condition:
s
(G(s
+
w
, T
+
) G(s
w
, T
)) = u
+
F(s
+
w
, T
+
) u
F(s
w
, T
) [Q], (5.35)
This term should be understood in the sense of distributions.
5.5.1 Water Evaporation Shock
WES - This is the discontinuity between a ( ) state (1,T
,u
) in the wr and a (+)
state (s
+
w
, T
b
,u
+
) in the br. It satisfies the following RH conditions for the speed v
WES
,
obtained from Eqs.
(5.8)-( 5.9):
u
+
(
ρ
w
f
+
w
+
ρ
b
g
f
+
g
)
ϕ
v
WES
(
ρ
w
s
+
w
+
ρ
b
g
s
+
g
)=u
ρ
w
ϕ
v
WES
ρ
w
, (5.36)
u
+
ρ
w
h
b
w
f
+
w
+
ρ
b
g
h
b
g
f
+
g
u
ρ
w
h
w
= v
WES
ϕ
ˆ
H
b
r
+
ρ
w
h
b
w
s
+
w
+
ρ
b
g
h
b
g
s
+
g
ˆ
H
r
ρ
w
h
w
.
(5.37)
111
where f
w
(s
w
= 1,·)=1, f
+
w
= f
w
(s
+
w
, T
b
) and f
g
= 1 f
w
.
The Darcy speed u
+
is found from u
using (5.36) and (5.37) as:
u
+
= u
(
ˆ
H
b
r
ˆ
H
r
)+s
+
w
ρ
w
h
b
w
h
w
+ s
+
g
ρ
b
g
h
b
g
h
w
(
ˆ
H
b
r
ˆ
H
r
)
f
+
g
ρ
b
g
/
ρ
w
+ f
+
w
+
ρ
b
g
(s
+
w
f
+
w
)+
ρ
w
f
+
w
h
b
w
h
w
+ s
+
g
ρ
b
g
h
b
g
h
w
.
(5.38)
Eq
(5.38) is always valid because T
< T
b
, so each term in the denominator is pos-
itive. The terms
ˆ
H
b
r
ˆ
H
r
and h
b
w
h
w
are positive because the enthalpies increase
with temperature. The term h
b
g
h
w
is positive because h
b
g
> h
b
w
and since h
b
w
> h
w
the positivity follows.
Since we can write u
+
as function of u
w
, we obtain v
WES
= v
WES
(T
,u
;s
+
w
):
v
WES
=
u
+
ϕ
f
+
g
ρ
b
g
h
b
g
h
w
+ f
+
w
ρ
w
h
b
w
h
w
ˆ
H
b
r
ˆ
H
r
+ s
+
g
ρ
b
g
h
b
g
h
w
+ s
+
w
ρ
w
h
b
w
h
w
. (5.39)
The denominator of v
WES
in (5.39) is never zero because each term in the sum is
positive.
Lemma 5.5.1. Define
L (T) as:
L (T)=
ρ
b
g
(h
b
g
h
T
w
)
ρ
w
h
b
w
h
T
w
, (5.40)
where h
T
w
= h
w
(T). There is a unique temperature T
< T
b
such that L(T
)=0.
Moreover,
L (T) < 0 for T < T
and L(T) > 0 for T > T
.
Proof: We define:
L
b
=
ρ
w
h
b
w
ρ
b
g
h
b
g
= H
b
w
H
b
g
,
where H
w
and H
g
are defined in 5.4. From (i)-(iii) of Remark 5.4, we can see that
L
b
> 0. Since L
b
does not depend on T,ifT is small, we obtain that
ρ
w
h
b
w
ρ
b
g
h
b
g
= L
b
> h
T
w
(
ρ
w
ρ
T
g
)=⇒L(T)=h
T
w
(
ρ
w
ρ
T
g
) −L
b
< 0,
On the other hand, if T
= T
b
we obtain from Eq. (5.40) that
L (T
b
)=L
b
> 0,
so there is almost a temperature T
such that L(T
)=0. Moreover, the temperature
T
is unique, because if we differentiate L (T):
d
L (T)
dT
= C
w
(
ρ
w
ρ
b
g
)/
ρ
w
> 0,
i.e,
L (T) is a monotonic function, so there is only a temperature that satisfies L(T
)=
0.
112
Substituting s
+
g
= 1 s
+
w
, f
+
g
= 1 f
+
w
, using Lemma 5.5.1 for T = T
, we rewrite
Eq.
(5.39)
v
WES
=
u
+
ϕ
f
+
w
f
WES
w
s
+
w
s
WES
w
, f
WES
w
ρ
b
g
(h
b
g
h
w
)
L (T
)
, s
WES
w
ˆ
H
b
r
ˆ
H
r
+
ρ
b
g
(h
b
g
h
w
)
L (T
)
,
(5.41)
Eqs.
(5.41) are the basis for a graphical construction of the WES, see Fig. 5.2.b.
For T
< T
, f
WES
w
and s
WES
w
are negative, while for T
> T
, f
WES
w
> 1 and
s
WES
w
> 1. When T
= T
, we obtain that
ρ
b
g
(h
b
g
h
w
)=
ρ
w
h
b
w
h
w
, so Eq.
(5.39)
reduces to:
v
WES
(1,T
= T
,u
w
;s
+
w
)=
u
+
ϕ
ρ
w
h
b
w
h
w

f
b
w
+ f
b
g
ˆ
H
b
r
H
r
+
ρ
w
(h
b
w
h
w
)
s
w
+ s
g
=
u
ϕ
C
w
ˆ
C
r
+ C
w
.
(5.42)
Notice that if u
= u
+
, v
WES
= v
w
T
in the wr, see Eq (5.32).
5.5.2 Vaporization Shock
VS - It is a discontinuity between a () state (s
w
, T
b
,u
) in the br and a (+) state
(0,T
+
> T
b
,u
+
) in the sr. The Vaporization Shock satisfies the following RH condi-
tions with speed v
VS
obtained from Eqs. (5.8)-(5.9):
v
VS
ϕ
(
ρ
+
g
ρ
b
g
s
g
ρ
w
s
w
)=u
+
ρ
+
g
u
(
ρ
w
f
w
+
ρ
b
g
f
g
), (5.43)
v
VS
ϕ
(H
+
r
H
b
r
+
ρ
+
g
h
+
g
s
g
ρ
b
g
h
b
g
s
w
ρ
w
h
b
w
)=u
+
ρ
+
g
h
+
g
u
(
ρ
b
g
h
b
g
f
g
+
ρ
w
h
b
w
f
w
),
(5.44)
where h
+
g
= h
g
(T
+
), h
+
w
= h
w
(T
+
), H
+
r
= H
r
(T
+
),
ρ
+
g
=
ρ
g
(T
+
) and f
w
= f
w
(s
w
, T
b
).
Since T
> T
b
, we obtain v
VS
as the following fraction, which has positive denomi-
nator:
v
VS
= v
VS
(s
w
,u
; T
+
)=
u
ϕ
f
g
ρ
b
g
h
+
g
h
b
g
+ f
w
ρ
w
h
+
g
h
b
w
ˆ
H
+
r
ˆ
H
b
r
+ s
g
ρ
b
g
h
+
g
h
b
g
+ s
w
ρ
w
h
+
g
h
b
w
. (5.45)
We rewrite Eq.
(5.45) in a shorter form. Substituting s
g
= 1 s
w
and f
g
= 1 f
g
,
we define
A(T
+
)=
ρ
b
g
h
+
g
h
b
g
ρ
w
h
+
g
h
b
w
.
Lemma 5.5.2. If T
+
> T
b
, the water and rock enthalpies satisfy the following in-
equality:
ρ
b
g
h
+
g
h
b
g
<
ρ
w
h
+
g
h
b
w
.
Proof: From Remark 5.1.1, we know that h
b
w
< h
b
g
. Using the expression for A we
obtain:
ρ
w
h
+
g
h
b
w
ρ
b
g
h
+
g
h
b
g
>
ρ
w
h
+
g
h
b
g
ρ
b
g
h
+
g
h
b
g
=(
ρ
w
ρ
b
g
)
h
+
g
h
g
,
113
but (
ρ
w
ρ
b
g
) > 0 because
ρ
w
>>
ρ
g
and (h
+
g
h
b
g
) > 0 since h
g
increases with
temperature.
From above Lemma 5.5.2, we obtain that A < 0. We multiply and divide (5.45) by
A and we obtain:
v
VS
=
u
ϕ
f
w
f
VS
w
s
w
s
VS
w
, where f
VS
w
ρ
w
h
+
g
h
b
w
A(T
+
)
, s
VS
w
ˆ
H
+
r
ˆ
H
b
r
+
ρ
b
g
h
+
g
h
b
g
A(T
+
)
;
(5.46)
notice that f
VS
w
and s
VS
w
are negative. Eqs. (5.46) are the basis for a graphical con-
struction of VS, see Fig. 5.2. a.
The Darcy speed u
+
in the sr is the fraction with positive denominator:
u
+
= u
A
+
(
ρ
w
(s
g
f
g
)+
ρ
+
g
f
g
)+B
+
(
ρ
b
g
(s
w
f
w
)+f
w
ρ
+
g
)+C
+
f
g
ρ
b
g
+ f
w
ρ
w
A
+
s
g
+ B
+
s
w
+ C
+
ρ
+
g
,
(5.47)
where:
A
+
=
ρ
b
g
h
+
g
h
b
g
, B
+
=
ρ
w
h
+
g
h
b
w
, C
+
=
ˆ
H
+
r
ˆ
H
b
r
. (5.48)
5.5.3 Condensation Shock
CS - This is the discontinuity between a () state (0,T
> T
b
,u
) in the sr and a
(+) state (s
+
w
, T
b
,u
+
) in the br. It is the reverse of the shock VS. From (5.47):
u
+
= u
A
s
+
g
+ B
s
+
w
+ C
ρ
g
A
(
ρ
w
(s
+
g
f
+
g
)+
ρ
g
f
+
g
)+B
(
ρ
b
g
(s
+
w
f
+
w
)+f
+
w
ρ
g
)+C
f
+
g
ρ
b
g
+ f
+
w
ρ
w
,
(5.49)
A
, B
and C
are obtained from A
+
, B
+
and C
+
in (5.48) by substituting T
+
by T
.
Lemma 5.5.3. The denominator of
(5.49) is always positive.
Proof: We calculate:
A
ρ
w
f
g
+ B
ρ
b
g
f
w
=
h
g
h
b
g
ρ
b
g
ρ
w
f
g
+
h
g
h
b
w
ρ
w
ρ
b
g
f
w
,
= h
g
ρ
b
g
ρ
w
f
g
h
b
g
ρ
w
f
g
+ h
g
ρ
w
ρ
b
g
f
w
h
b
w
ρ
b
g
f
w
,
= h
g
ρ
b
g
ρ
w
( f
g
+ f
w
) (h
b
g
ρ
w
f
g
+ h
b
w
ρ
b
g
f
w
),
= h
g
ρ
b
g
ρ
w
(h
b
g
ρ
w
f
g
+ h
b
w
ρ
b
g
f
w
), (5.50)
where A
and B
are given in (5.48) and f
w
= f
+
w
and f
g
= f
+
g
.
Similarly, we calculate:
A
ρ
w
s
g
+ B
ρ
b
g
s
w
= h
g
ρ
b
g
ρ
w
(h
b
g
ρ
w
s
g
+ h
b
w
ρ
b
g
s
w
), (5.51)
114
where s
w
= s
+
w
and s
g
= s
g
. So, from (5.51) and (5.50):
A
ρ
w
(s
g
f
g
)+B
ρ
b
g
(s
w
f
w
)=h
b
g
ρ
w
( f
g
s
g
)+h
b
w
ρ
b
g
( f
w
s
w
),
=
ρ
w
ρ
b
g
h
b
g
( f
g
s
g
)+h
b
w
( f
w
s
w
)
,
=
ρ
w
ρ
b
g
h
b
g
h
b
w
(s
w
f
w
), (5.52)
where
ρ
w
ρ
b
g
h
b
g
h
b
w
> 0.
We rewrite the denominator of
(5.49) as:
C
f
b
g
ρ
b
g
+ f
b
w
ρ
w
+
ρ
w
ρ
b
g
h
b
g
h
b
w
(s
w
f
w
)+A
ρ
g
f
g
+ B
ρ
g
f
w
; (5.53)
notice that
B
ρ
g
ρ
w
ρ
b
g
h
b
g
h
b
w
=
ρ
w
H
g
ρ
g
H
b
w
ρ
w
H
b
g
ρ
b
g
H
b
w
, (5.54)
where H
g
and H
w
are defined in 5.1.1. From the same Remark, we know that H
g
> H
b
g
because T
> T
b
, (see (5.54)) and
ρ
b
g
>
ρ
g
because
ρ
g
is decreasing, we have:
B
ρ
g
ρ
w
ρ
b
g
h
b
g
h
b
w
>
ρ
w
H
b
g
ρ
b
g
H
b
w
ρ
w
H
b
g
ρ
b
g
H
b
w
= 0, (5.55)
so from
(5.53) and (5.55) we obtain:
C
f
b
g
ρ
b
g
+ f
b
w
ρ
w
+
ρ
w
ρ
b
g
h
b
g
h
b
w
(s
w
f
w
)+A
ρ
g
f
g
+ B
ρ
g
f
w
,
= C
f
b
g
ρ
b
g
+ f
b
w
ρ
w
+
ρ
w
ρ
b
g
h
b
g
h
b
w
s
w
+ A
ρ
g
f
g
+ f
w
(B
ρ
g
ρ
w
ρ
b
g
h
b
g
h
b
w
),
> C
f
b
g
ρ
b
g
+ f
b
w
ρ
w
+
ρ
w
ρ
b
g
h
b
g
h
b
w
s
w
+ A
ρ
g
f
g
> 0. (5.56)
The last equality is because h
b
g
> h
b
w
and each term is positive. So the denominator of
(5.49) is always positive.
Since the CS is reverse of the VS and u
+
is a function of u
, we obtain v
CS
=
v
CS
(T
,u
;s
+
w
):
v
CS
=
u
+
ϕ
f
+
g
ρ
b
g
h
g
h
b
g
+ f
+
w
ρ
w
h
g
h
b
w
ˆ
H
r
ˆ
H
b
r
+ s
+
g
ρ
b
g
h
g
h
b
g
+ s
+
w
ρ
w
h
g
h
b
w
or v
CS
=
u
+
ϕ
f
+
w
f
CS
w
s
+
w
s
CS
w
.
(5.57)
Since the CS shock is reverse of the VS, f
CS
w
and s
CS
w
are obtained from f
VS
w
and s
VS
w
in
(5.46) by substituting T
+
by T
; notice that the denominator in (5.57.a) is never zero.
115
5.5.4 Steam condensation front
SCF - This is the discontinuity between a () state (s
w
< 1,T
b
,u
) in the br and a
(+) state (1,T
+
,u
+
) in the wr. It is the reverse of the WES, so v
SC F
= v
SC F
(s
w
,u
; T
+
)
is given by:
v
SC F
=
u
ϕ
f
g
ρ
b
g
h
b
g
h
+
w
+ f
w
ρ
w
h
b
w
h
+
w
ˆ
H
b
r
ˆ
H
+
r
+ s
g
ρ
b
g
h
b
g
h
+
w
+ s
w
ρ
w
h
b
w
h
+
w
=
u
ϕ
f
w
f
SC F
w
s
w
s
SC F
w
. (5.58)
Here
(5.58) is derived as Eq. (5.41), with f
SC F
w
and s
SC F
w
obtained from f
WES
w
and s
WES
w
in (5.41) by substituting T
+
by T
.
From Eqs.
(5.36)-( 5.37), we can find u
+
as the following fraction with positive
denominator:
u
+
= u
(
ˆ
H
b
r
ˆ
H
+
r
)
f
g
ρ
b
g
/
ρ
w
+ f
w
+
ρ
b
g
(s
w
f
w
)+
ρ
w
f
w
h
b
w
h
+
w
+ s
g
ρ
b
g
h
b
g
h
+
w
ˆ
H
b
r
ˆ
H
+
r
+ s
w
ρ
w
h
b
w
h
+
w
+ s
g
ρ
b
g
h
b
g
h
+
w
.
(5.59)
5.6 The Riemann Solution
The Riemann problem is the solution of (5.8)-(5.9) with initial data
L
=(s
L
, T
L
,u
L
) if x > 0
R
=(s
R
, T
R
,·) if x < 0,
(5.60)
where s :
= s
w
is the water saturation. We will see that the speed u cannot be pre-
scribed on both sides. Given a speed on one side the other one is obtained by solving
the system
(5.1)-( 5.3); in this case we have chosen to prescribe u
L
.
We consider in this paper the Riemann problem for all initial data; we divide the
data as:
Riemann Problem L state R state
Data A Steam and Water, T
L
= T
b
Steam, T
R
> T
b
Data B Water, T
L
< T
b
Steam and Water, T
R
= T
b
Data C Steam, T
L
> T
b
Steam and Water, T
R
= T
b
Data D Water, T
L
< T
b
Steam, T
R
> T
b
Data E Steam, T
L
> T
b
Water, T
R
< T
b
Data F Steam and water, T
L
= T
b
Water, T
R
< T
b
The Riemann problem with Data F for T
R
= T
0
was solved in [6]; in that paper,
T
0
is an arbitrary temperature where the water and rock enthalpies were made to
vanish; we do not repeat this Riemann problem here.
The rarefaction waves are denoted by R
T
for thermal rarefactions in sr and R
s
for
(Buckley-Leverett) saturation rarefactions in br. The shocks in regions under ther-
modynamic equilibrium are denoted with a single subscript, S
T
for thermal shocks in
116
sr, S
s
for (Buckley-Leverett) saturation shocks , S
G
for HIWS in br and S
W
for thermal
discontinuities in wr. We recall that the shocks between regions are WES, VS, CS and
SC F.
5.6.1 Riemann Problem A
Water injection.
First, we inject water with temperature T
b
, i.e., L =(1,T
b
,u
L
) in a porous rock filled
with superheated steam, i.e., R
=(0,T
R
> T
b
,u
R
), which is a sr. In the br, generated
by the L state, the flow is governed by a Buckley-Leverett equation. It is well known
that the Buckley-Leverett rarefaction has speed
λ
b
s
given by (5.30) from s
w
= 1 to
s
w
= s
, which is defined by:
f
w
s
w
(s
, T
b
)=
f
w
(s
, T
b
)
s
. (5.61)
There is a saturation shock and the solution is continued by a shock in the br.
Since the temperature increases in the rarefaction joining the
() state (0,T
b
,u
)
to the (+) state (0,T
+
> T
b
,u
+
), see Rem. 5.4.3, there is a shock with speed v
g
T
given
by
(5.25).
Lemma 5.6.1. The speed v
b
g,s
of the HIWS shock between (s
, T
b
,u
) and (0,T
b
,u =
u
) given by (5.29) is larger than the speed v
g
T
of the shock between (0, T
b
,u
) and
(0,T > T
b
,u
+
).
Proof: From Eq.
(5.25) we have u
= u
b
and we can see easily that v
g
T
< u
b
/
ϕ
.
From
(5.29) and (5.61), we have that v
b
g,s
for s
satisfies v
b
g,s
> u
b
/
ϕ
. So we conclude
that v
g
T
< v
b
g,s
.
Thus we conclude that there is a shock with speed v
VS
between the br and the sr.
Let f
VS
w
and s
VS
w
be given by (5.46); from the Sec. 2.1.3, we find a saturation
ˆ
s > s
defined by the following equality, (Fig. 5.2.a):
v
VS
(
ˆ
s
)=
u
b
ϕ
f
w
(
ˆ
s, T
b
) f
VS
w
ˆ
s
s
VS
w
=
u
b
ϕ
f
w
s
w
(
ˆ
s, T
b
). (5.62)
Lemma 5.6.2. The saturations
ˆ
s and s
satisfy s
<
ˆ
s
Proof: The result follows from shape of f
w
, i.e., a function convex between s
w
= 0
and s
inf
; and concave between s
inf
and s
w
= 1 with f
w
(1,·)=1.
Using Eq.
(5.62) we can prove the following Lemma:
Lemma 5.6.3. The solutions of
(5.62) are the points where the derivative of v
VS
with
respect to S
w
vanishes.
117
Proof: The points that satisfy (5.62) are tangency points of the secant from the
point
(S
VS
w
, f
VS
w
) to the graph f
w
, so the derivative vanishes. (An analytic proof for this
Lemma is similar to proof of the Lemma 5.6.6).
Moreover, we can verify the following Proposition:
Proposition 5.6.1. There are two saturation values that satisfy
(5.62) for each T
+
>
T
b
. The largest value satisfies (5.62) and maximizes v
VS
while the other minimizes
v
VS
; because of geometrical compatibility, we choose the largest value and we denote it
by
ˆ
s. It is called “Hot-Bifurcation saturation I” or HBI.
Proof: We verify geometrically this proposition; equation
(5.62) represents tan-
gency points of the secant from the point
(S
VS
w
, f
VS
w
) to the graph f
w
, for each fixed T
.
In figure 5.2, we plot a possible point
(S
VS
w
, f
VS
w
).
The largest water saturation value satisfying
(5.62) maximizes v
VS
. This value is
denoted
ˆ
S, and the other saturation minimizes the shock speed v
VS
.
The analytical proof for this proposition can be obtained following the steps of
Proposition 5.6.4, however this proof is longer than the geometrical verification. No-
tice that in Proposition 5.6.4, the chosen water saturation minimizes v
CS
.
Figure 5.2: The coincidence between v
VS
and
λ
b
S
.AsA < 0,soS
VS
w
< 0 and f
VS
w
< 0
are negative, the largest water saturation point that satisfies
(5.62) maximizes and
the other point minimizes v
VS
.
It is necessary that v
VS
(
ˆ
s
) > v
g
T
; otherwise, the geometrical compatibility in Sec.
2.1.3 says that the vaporization shock VS does not exists. This is summarized as
follows:
Lemma 5.6.4. The vaporization shock between
() state (s
w
, T
b
,u
) and (+) state
(1,T
+
,u
+
) with speed v
VS
given by (5.45) is larger than v
g
T
for s
s
w
1 and
T
+
> T
b
. (We recall that s
is defined in Eq. (5.61)).
118
Proof: From (5.45), v
VS
is written as:
v
VS
(S
w
)=
u
b
ϕ
f
g
ρ
b
g
h
+
g
h
b
g
+ f
w
ρ
w
h
+
g
h
b
w
ˆ
H
+
r
ˆ
H
r
+ s
g
ρ
b
g
h
+
g
h
b
g
+ s
w
ρ
w
h
+
g
h
b
w
, (5.63)
where s
w
= s
w
and s
g
= s
g
.Astos
s
w
1,wehavef
w
(s
w
) > s
w
, so from Lemma
5.5.2 and
(5.63) we obtain:
v
VS
(S
w
) >
u
b
ϕ
f
g
ρ
b
g
h
+
g
h
b
g
+ f
w
ρ
b
g
h
+
g
h
b
g
ˆ
H
+
r
ˆ
H
r
+ s
g
ρ
b
g
h
+
g
h
b
g
+ s
w
ρ
b
g
h
+
g
h
b
g
; (5.64)
using f
w
+ f
g
= 1 and S
w
+ S
g
= 1 in (5.64) we write finally:
v
VS
(S
w
) >
u
b
ϕ
ρ
b
g
h
+
g
h
b
g
ˆ
H
+
r
ˆ
H
r
+
ρ
b
g
h
+
g
h
b
g
= v
g
T
. (5.65)
From
(5.29), we conclude that v
b
g,s
and s
w
decrease together. From (5.25), v
g
T
is
constant. So we expect that there is a saturation s
∗∗
where v
g
T
= v
b
g,s
(s
∗∗
,s
w
= 0),
which is called “Hot-Bifurcation II saturation”, or HBII.
Remark 5.6.1. Let a,b, c, d
R
,if
a
b
=
c
d
,
−→
a
b
=
c
α
a
d
α
b
=
a
α
c
b
α
d
=
c
d
, for
α
= k.
One can verify the following:
Proposition 5.6.2. For each fixed T
+
, there is a unique saturation s
∗∗
for which
v
g
T
(T
b
,u
; T
+
)=v
b
g,s
(u
;s
∗∗
)=v
VS
(s
∗∗
,u
; T
+
). (5.66)
Furthermore s
∗∗
is defined in terms of T
+
by any of the following equivalent equalities:
1. v
g
T
(T
b
,u
; T
+
)=v
b
g,s
(u
;s
∗∗
);
2. v
b
g,s
(u
;s
∗∗
)=v
VS
(s
∗∗
,u
; T
+
);
3. v
g
T
(T
b
,u
; T
+
)=v
VS
(s
∗∗
,u
; T
+
).
Proof:
(1) First we assume that v
g
T
= v
b
g,s
:
We set T
b
= T
and T = T
+
, so we can write
v
g
T
=
u
ϕ
ρ
g
(h
+
g
h
g
)
ˆ
H
+
r
ˆ
H
r
+
ρ
g
(h
+
g
h
g
)
=
u
ϕ
f
w
S
∗∗
= v
b
g,s
, (5.67)
119
where f
w
= f
w
(s
∗∗
, T
b
).
We multiply and divide the third term of
(5.67) by
ρ
w
h
+
g
h
w
, which is positive:
v
g
T
=
u
ϕ
ρ
g
(h
+
g
h
g
)
ˆ
H
+
r
ˆ
H
r
+
ρ
g
(h
+
g
h
g
)
=
u
ϕ
f
w
ρ
w
h
+
g
h
w
s
∗∗
ρ
w
h
+
g
h
w
= v
b
g,s
. (5.68)
From the Remark 5.6.1, we obtain:
v
g
T
=
u
ϕ
ρ
g
(h
+
g
h
g
)+f
w
ρ
w
h
+
g
h
w
ˆ
H
+
r
ˆ
H
r
+
ρ
g
(h
+
g
h
g
)+s
∗∗
ρ
w
h
+
g
h
w
=
u
ϕ
f
w
s
∗∗
= v
b
g,s
. (5.69)
We multiply and divide the third equation of
(5.69) by
ρ
w
h
+
g
h
w
,sov
g
T
equals:
u
ϕ
ρ
g
(h
+
g
h
g
)+f
w
ρ
w
h
+
g
h
w
ˆ
H
+
r
ˆ
H
r
+
ρ
g
(h
+
g
h
g
)+s
∗∗
ρ
w
h
+
g
h
w
=
u
ϕ
f
w
ρ
g
(h
+
g
h
g
)
s
∗∗
ρ
g
(h
+
g
h
g
)
=
v
b
g,s
. (5.70)
From the Remark 5.6.1, and using that 1
f
w
(s
∗∗
)= f
g
and 1 s
∗∗
= s
∗∗
g
we obtain:
v
g
T
=
u
ϕ
f
g
ρ
g
(h
+
g
h
g
)+f
w
ρ
w
h
+
g
h
w
ˆ
H
+
r
ˆ
H
r
+ s
∗∗
g
ρ
g
(h
+
g
h
g
)+s
∗∗
ρ
w
h
+
g
h
w
= v
VS
= v
b
g,s
. (5.71)
so v
g
T
= v
b
g,s
implies that v
g
T
= v
b
g,s
= v
VS
.
(2) We assume that v
b
g,s
= v
VS
so:
v
VS
=
u
ϕ
f
g
ρ
g
(h
+
g
h
g
)+f
w
ρ
w
h
+
g
h
w
ˆ
H
+
r
ˆ
H
r
+ s
∗∗
g
ρ
g
(h
+
g
h
g
)+s
∗∗
ρ
w
h
+
g
h
w
=
u
ϕ
f
w
s
∗∗
= v
b
g,s
(5.72)
Performing the above calculation in reverse order, we see that v
VS
= v
b
g,s
implies that
v
VS
= v
b
g,s
= v
g
T
.
(3) We assume that v
g
T
= v
VS
, so:
ρ
g
(h
+
g
h
g
)
ˆ
H
+
r
ˆ
H
r
+
ρ
g
(h
+
g
h
g
)
=
u
ϕ
f
g
ρ
g
(h
+
g
h
g
)+f
w
ρ
w
h
+
g
h
w
ˆ
H
+
r
ˆ
H
r
+ s
∗∗
g
ρ
g
(h
+
g
h
g
)+s
∗∗
ρ
w
h
+
g
h
w
. (5.73)
From the Remark 5.6.1, we obtain:
ρ
g
(h
+
g
h
g
)
ˆ
H
+
r
ˆ
H
r
+
ρ
g
(h
+
g
h
g
)
=
f
w
ρ
w
h
+
g
h
w
ρ
g
(h
+
g
h
g
)
s
∗∗
ρ
w
h
+
g
h
w
ρ
g
(h
+
g
h
g
)
. (5.74)
120
Since
ρ
w
h
+
g
h
w
ρ
g
(h
+
g
h
g
) = 0 (see Lemma 5.5.2), then:
ρ
g
(h
+
g
h
g
)
ˆ
H
+
r
ˆ
H
r
+
ρ
g
(h
+
g
h
g
)
=
f
w
s
∗∗
. (5.75)
so v
g
T
= v
VS
implies that v
g
T
= v
b
g,s
= v
VS
.
Lemma 5.6.5. For any T
+
> T
b
in the physical range, the saturation s
∗∗
given by Eq.
(5.66) satisfies s
∗∗
<
ˆ
s.
Proof: The saturation s
∗∗
satisfies Eq. (1) of Proposition 5.6.2. Geometrically,
this Equation represent a line from a point
(s
VS
w
, f
VS
w
) passing on the point (0,0)
and intersecting the graph f
w
for each fixed T
+
, where s
VS
w
and f
VS
w
are given by Eq.
(5.46.c) and (5.46.b) respectively. Notice that this intersection occurs always because
s
VS
w
< f
VS
w
< 0. Moreover, from Figure 5.3 we can notice that s
∗∗
ˆ
s and s
∗∗
=
ˆ
s if,
only if, s
= s
∗∗
=
ˆ
s, and from Lemma 5.6.2 follows that s
<
ˆ
s,sos
∗∗
<
ˆ
s. Since T
+
is
arbitrary the Lemma is proved.
**
Figure 5.3: The saturation s
∗∗
obtained from a line from a point (s
VS
w
, f
VS
w
) passing on
the point
(0,0) and intersecting the graph f
w
.
Solution. Now we can describe the possible solutions for Riemann Data A:
For
ˆ
s
s
L
1. The waves R
s
VS, with
ˆ
s given by Eq. (5.62) and the sequence:
L
=(s
L
, T
b
,u
L
)
R
s
(
ˆ
s, T
b
,u
L
)
VS
−→ (0,T
R
,u
R
)=R. (5.76)
For s
∗∗
s
L
<
ˆ
s. The wave VS with s
∗∗
given by Eq. (5.66) and the sequence:
L
=(s
L
, T
b
,u
L
)
VS
−→ (0,T
R
,u
R
)=R. (5.77)
121
For s
L
< s
∗∗
. The waves S
G
S
T
, with sequence:
(s
L
, T
b
,u
L
)
S
G
(0,T
b
,u
L
)
S
T
(0,T
R
,u
R
). (5.78)
Wave behavior on bifurcations
Notice that the Buckley-Leverett rarefaction wave R
S
disappears on the vertical dot-
ted line from
ˆ
S that separates region I from II; we call it the R
S
S
VS
boundary
bifurcation”. On the vertical dotted line from S
∗∗
separating region II from III, the
condensation shock S
VS
between regions disappears and a saturation S
G
and a ther-
mal shock S
T
appear in the br; we call this line the S
T
S
VS
bifurcation”.
Figure 5.4: The dotted line is the boundary of physical range. The dashed lines are
bifurcation loci. The temperature is higher than the water boiling temperature (su-
perheated steam) or is equal to boiling temperature. The saturations S
∗∗
and
ˆ
S are
given in Eqs.
(5.66) and (5.62); the vertical lines from
ˆ
S and S
∗∗
are the R
S
S
VS
bifurcation” and the S
T
S
VS
bifurcation”.
5.6.2 Riemann Problem B
Since the flow in the wr is governed by the linear Eq. (5.31) with constant character-
istic speed v
w
T
given by (5.32), this wave is a contact discontinuity.
For the Riemann problem L
=(1,T
L
T
b
,u
L
) and R =(s
R
, T
b
,u
R
), we have the
following:
Proposition 5.6.3. For each
() state (1,T
< T
b
,u
) and (+) state (s
+
w
, T
+
,u
+
),
there is a water saturation s
= s
(T
), such that for s
+
w
satisfying s
s
+
w
1, the
shock speeds v
WES
and v
w
T
satisfy:
v
w
T
v
WES
, s
s
+
w
1; (5.79)
122
the equality occurs only if s
+
w
= 1 or s
+
w
= s
. We call s
a “Cold Bifurcation saturation
I” or CBI. Moreover, there is a water saturation s
††
= s
††
(T
) satisfying s
< s
††
such
that:
λ
b
s
(s
††
)=v
WES
(T
,u
;s
††
). (5.80)
Proof: We define the functions Θ
(T
;s
+
w
) and Ψ(T
;s
+
w
) as:
Θ
(T
;s
+
w
) := v
w
T
v
WES
(T
;s
+
w
), (5.81)
Ψ
(T
;s
+
w
) :=
λ
b
s
(T
;s
+
w
) v
WES
(T
;s
+
w
). (5.82)
The proposition is true, because from Figure 5.5 the curve Ψ is above the curve Θ.
300 310 320 330 340 350 360 370
0.98
0.982
0.984
0.986
0.988
0.99
0.992
0.994
0.996
0.998
1
Temperature/K in the Water Region / T
W
ater Saturation in the Boiling Region / s
w
+
Ψ
Θ
Figure 5.5: The functions Θ(T
;s
+
w
) and Ψ(T
;s
+
w
) in the {T
,s
+
w
}.
In the nomenclature of
[26], the state s
w
= 1 is the left-extension of s
††
with speed
λ
b
s
. Also, s
††
here coincides with s
††
obtained in [6]. This saturation maximizes v
WES
(and consequently v
SC F
); we call s
††
the “Cold Bifurcation saturation II” or CBII.
Remark 5.6.2. Notice that from Sec. 5.5.1, f
WES
w
and s
WES
w
are negative if T
< T
and
positive if T
> T
(see Fig. 5.2.b). The solution behavior is the same in both cases.
Remark 5.6.3. Notice that from Section 5.5.1, f
WES
w
and S
WES
w
are negative if T < T
and positive if T > T
(see Figure 5.6). Nevertheless the solution behavior is the same,
i.e., the point T
is not a bifurcation.
Prop. 5.6.3 yields the following Corollaries used to obtain the solution in the br:
123
Figure 5.6: We represent two possible points s
WES
w
, f
WES
w
).IfT < T
, both s
WES
w
and
f
WES
w
are negative; otherwise both are larger than 1.
Corollary 5.6.1. If s
infl
(T
b
) < s
+
w
< s
††
, the solution continues in the br as a rarefac-
tion to s
+
w
.Ifs
+
w
< s
infl
the rarefaction continues to s
§
, where s
§
is defined by the second
equality in:
v
§,+
=
f
w
s
w
(s
§
, T
b
)=
f
w
(s
+
w
, T
b
) f
w
(s
§
, T
b
)
s
+
w
s
§
, (5.83)
where s
infl
= s
infl
(T
b
) is the inflection saturation defined by:
2
s
2
w
f
w
(s
w
, T
b
)
s
w
=s
infl
= 0. (5.84)
Proof: From Section 5.4.3 and
[6 ], we know that v
w
T
satisfies:
v
w
T
ϕ
u
w
=
C
w
ˆ
C
r
+ C
w
< 1,
the last inequality holds because
ˆ
C
r
and C
w
are positive, and from Eq. (5.79) in Propo-
sition 5.6.3 we obtain that v
WES
< 1.
If s
w
> s
where s
satisfies (5.61), the rarefaction wave is always a solution but at
s
we know that:
f
w
(
s
)
s
w
> 1.
As v
WES
satisfy (5.80) we obtain that:
f
w
(s
††
, T
b
)
s
w
<
f
w
(s
, T
b
)
s
w
,
124
so if s
infl
< s
+
w
there is a rarefaction from s
††
to s
+
w
and if s
+
w
< s
infl
there is a rarefac-
tion from s
††
to s
§
that satisfies (5.83).
Corollary 5.6.2. As the left state temperature T
tends to the water boiling tempera-
ture, the water saturation s
††
tends to 1, i.e., the limit of s
††
lies in the wr.
Proof: We take the limit T
−→ T
b
in Eq. (5.38), and obtain:
u
b
= u
w
s
g
ρ
b
g
(h
b
g
h
g
w
)
s
g
ρ
b
g
(h
b
g
h
g
w
)
=
u
w
,
because h
b
g
> h
g
w
. Substituting u
b
= u
w
in Eq. (5.39) and taking the limit T
−→ T
b
:
v
WES
=
u
b
ϕ
f
b
g
S
g
.
From Eq.
(5.80), after simplifications, we obtain that at (s
††
, T):
f
w
s
w
=
1 f
w
1 s
††
, (5.85)
Eq.
(5.85) has two solutions, one of which is s
††
= 1. The other solution s
††,2
satisfies
s
††,2
< s
infl
, so only s
††
= 1 is important for the solution.
Solution. Now we can describe the possible solutions for Riemann Data B:
For s
R
> s
††
. As s
††
satisfies (5.80), (u
b
/
ϕ
)f
w
(s
R
, T
b
)/s
w
< v
WES
, i.e, the shock
v
WES
is faster than the characteristic speed in the br, the wave sequence is:
L
=(1,T
L
,u
L
)
WES
−−→ (s
R
, T
b
,u
R
)=R. (5.86)
For s
infl
(T
b
) < s
R
< s
††
. The waves WES R
s
, with sequence:
L
=(1,T
L
,u
L
)
WES
−−→ (s
††
, T
b
,u
R
)
R
s
(s
R
, T
b
,u
R
)=R. (5.87)
For s
R
< s
infl
(T
b
). The waves WES R
s
S
s
with sequence:
L
=(1,T
L
,u
L
)
WES
−−→ (s
††
, T
b
,u
R
)
R
s
(s
§
, T
b
,u
R
)
S
s
(s
R
, T
b
,u
R
)=R, (5.88)
where s
§
is given by Eq. (5.83).
Wave behavior on bifurcations
The Buckley-Leverett rarefaction wave R
S
appears on the horizontal dotted line from
s
††
that separates region I from II; we call this line the S
WES
R
S
bifurcation”. On
the horizontal dotted line from S
infl
separating the region II from III, a Buckley-
Leverett shock S
S
appears; we call this line the R
S
S
S
bifurcation”.
125
Figure 5.7: The dotted line is the boundary of physical range. The dashed lines divides
are the bifurcation loci. The temperature is lower than the water boiling temperature
(liquid water) or is equal to water boiling temperature. The saturations s
††
and s
infl
are given in Eqs. (5.66) and (5.62); the horizontal lines from s
††
and s
infl
are the
S
WES
R
S
bifurcation” and the R
S
S
S
bifurcation”.
5.6.3 Riemann Problem C
Let L =(0,T
L
> T
b
,u
L
) and R =(s
R
, T
b
,u
R
). In this Riemann Problem, there are
two relevant bifurcation curves. The first bifurcation occurs at the points where the
thermal rarefaction speed
λ
g
T
(Eq. (5.17.a)) equals the shock speed v
CS
(Eq. (5.57)).
The other bifurcation appears at the points where the speed v
CS
coincides with the
Buckley-Leverett rarefaction speed
λ
b
s
(Eq. (5.30)) in the br.
Moreover, there is a point where all previous speeds coincide. It is a double bi-
furcation point in the
{T
;s
+
w
} plane of left state temperatures T
in sr and right
saturation s
+
w
in br (see Sec. 5.6.3). This point is denoted by (
ˆ
T;s
) and it should be
understood as the projection of
(s
w
= 0,
ˆ
T;s
, T
b
) onto the {T
;s
+
w
} plane.
Definition of
ˆ
T and s
Let Υ = Υ(T
;s
+
w
) be defined as:
Υ
=
ρ
g
c
g
ˆ
C
r
+
ρ
g
c
g
u
+
u
f
+
w
f
CS
w
s
+
w
s
CS
w
. (5.89)
The fraction u
+
/u
is obtained from Eq. (5.47). This fraction does not depend on u
or u
+
, showing that also Υ does not depend on u
(or u
+
). Moreover, at the points
(T
;s
+
w
) where Υ = 0, the equality
λ
g
T
= v
CS
holds.
126
Now we can define the thermal coincidence as the curve where the left thermal
wave speed
λ
g
T
coincides with the condensation shock v
CS
; it is denoted by TCS locus:
TCS
= {(T, s) | Υ(T, s)=0, for T sr and s br }. (5.90)
Analogously, we define Λ
= Λ(T
;s
+
w
) as:
Λ
=
f
+
w
f
CS
w
s
+
w
s
CS
w
f
b
w
s
w
, (5.91)
where f
CS
w
and s
CS
w
depend on T
, see (5.46) . Now we can define the CSS locus as the
curve where v
CS
=
λ
b
s
and for each T
fixed, v
CS
(T
;s
+
w
) is understood as a function
of s
+
w
that is minimized (see Prop. 5.6.4):
CSS
= {(T
;s
+
w
) | Λ = 0 | v
CS
is minimum; T
sr, s
+
w
br}. (5.92)
In Fig. 5.8, the TCS and CSS loci are shown as curves in the plane
{T,s}. The
horizontal axis represents the states in the sr and the vertical axis represents the
states in the br. The two loci intersect transversally at
(
ˆ
T;s
), the double bifurcation
point “SHB”. It can be obtained numerically using root finders. The temperature
ˆ
T
satisfies T
b
<
ˆ
T, and the saturation s
satisfies 0 < s
< 1.
For the Riemann solution, we need to study the relationships between T
L
and
ˆ
T at
(s
w
= 0,T
), and between s
R
and s
at (s
+
w
, T
b
) .
Let us define
Ξ
(T
;s
+
w
) :=
f
+
w
f
CS
w
s
+
w
s
CS
w
:=
f
+
g
ρ
b
g
h
g
h
b
g
+ f
+
w
ρ
w
h
g
h
b
w
ˆ
H
r
ˆ
H
b
r
+ s
+
g
ρ
b
g
h
g
h
b
g
+ s
+
w
ρ
w
h
g
h
b
w
(5.93)
the second equality originates from the definition f
CS
w
and s
CS
w
given in Eq. (5.41).
Lemma 5.6.6. The solutions of
(5.92) are the points where Ξ/s
w
is zero:
Proof: We differentiate
(5.93) with respect to s
w
, equate it to zero and isolate
f
w
/s
w
to obtain the result. We notice that f
g
/s
w
= f
w
/s
w
and s
g
/s
w
= 1.
Lemma 5.6.7. For all 0 s
w
1 the following inequality is valid
f
w
s
w
f
w
(s
infl
,·)
s
w
(5.94)
Proof: The extremum for f
w
/s
w
is obtained for /s
w
( f
w
/s
w
)=0; performing this
differentiation, we obtain that f
w
/s
w
= f
w
/s
w
, which yields
f
w
s
w
max
s
w
f
w
s
w
=
f
w
(s
infl
,·)
s
w
.
The CSS locus is obtained using the following proposition:
127
Figure 5.8: Schematic phase space. The intersection of the TCS and CSS loci is the
SHB at
(
ˆ
T;s
). The horizontal axis represents the sr and the vertical axis represents
the br,soSHB represents two points:
ˆ
T is the projection of
(s
w
= 0;
ˆ
T) on the sr;
s
is the projection of (s
w
= s
; T
b
) on the br; between the TCS locus and the water
saturation axis Υ
> 0,so
λ
g
T
> v
CS
; in the complementary region
λ
g
T
< v
CS
.
Proposition 5.6.4. There are always two water saturation values satisfying Λ
(T,s)=
0 for each fixed T > T
b
. The smallest value minimizes Ξ (and consequently v
CS
), while
the other value maximizes Ξ (and v
CS
). We define s
as the smaller water saturation
value.
The point
(T,s
) belongs to CSS locus.
Proof: The function f
w
/s
w
is continuous and satisfies the following relationship,
where f
w
/s
w
is evaluated at T = T
b
f
w
(s
w
= 0)
s
w
=
f
w
(s
w
= 1)
s
w
= 0. (5.95)
From equation
(5.93) with s
w
= 0 it follows that:
0
<
ρ
b
g
h
g
h
b
g
ˆ
H
r
ˆ
H
b
r
+
ρ
b
g
h
g
h
b
g
< 1. (5.96)
From equation
(5.93) with s
w
= 1 we have:
0
<
ρ
w
h
g
h
b
w
ˆ
H
r
ˆ
H
b
r
+
ρ
w
h
g
h
b
w
< 1. (5.97)
128
From Lemma 5.5.2 we know that
ρ
w
(h
+
g
h
b
w
>
ρ
b
g
(h
+
g
h
b
g
), from the fact that f
w
> s
w
we obtain that f
g
< s
g
and from Lemma 5.6.7 it follows that:
f
+
g
ρ
b
g
h
g
h
b
g
+ f
+
w
ρ
w
h
g
h
b
w
ˆ
H
r
ˆ
H
b
r
+ s
+
g
ρ
b
g
h
g
h
b
g
+ s
+
w
ρ
w
h
g
h
b
w
<
<
f
w
(s
infl
)
s
w
ρ
w
(h
+
g
h
b
w
)
ˆ
H
T
r
ˆ
H
b
r
+
ρ
w
(h
+
g
h
b
w
)
<
f
w
(s
infl
)
s
w
, (5.98)
so from the mean value theorem, for each fixed T, there is at least one point between
s
w
= 0 and s
w
= s
infl
and at least another point between s
w
= s
infl
and s
w
= 1 where
Ξ
(T;s) vanishes.
We need to show that there exist exactly 2 saturations that satisfy Ξ
(T;s)=0.To
do so we note that:
2
f
w
(s
w
)
s
2
w
= 0 ⇐⇒ s
w
= s
infl
. (5.99)
Moreover
2
f
w
/s
2
w
satisfies:
2
f
w
(s
w
)
s
2
w
> 0, if s
w
< s
infl
2
f
w
(s
w
)
s
2
w
< 0, if s
w
> s
infl
,
(5.100)
but from Lemma 5.6.6 we know that if s
w
satisfies Ξ(T, s)=0, then s
w
is a extremum
for
(5.93), i.e,
Ξ
=
f
w
s
w
⇐⇒
Ξ
s
w
= 0, (5.101)
where Ξ is the function defined in
(5.93).
Claim 1: The smaller water saturation for which Ξ
(T,s) vanishes is a minimum
for Ξ
(T,s).
Proof: We differentiate Ξ with respect to s
w
and obtain after manipulations that:
Ξ
s
w
=
B A
D
2
f
w
s
w
Ξ
, (5.102)
where A and B are given in
(5.48) and D is the denominator of Ξ. The term D satisfies
D
= 0, and from Lemma 5.5.2 we know that B A > 0, so the sign of (5.102) depends
only on the sign of:
=
f
w
s
w
Ξ, (5.103)
where is a continuous function.
From
(5.95) and (5.96), we know that f
w
/s
w
Ξ < 0 for s
w
= 0 or 1.IfΞ(T,s)=0
is satisfied we obtain that Ξ
/s
w
= 0 and then = 0; on the other hand if = 0 this
point satisfies Ξ
(T,s)=0 and Ξ/s
w
= 0.
129
We know that there are at least 2 water saturations where (5.92) is satisfied. We
denote the smaller one by s
, so we obtain that < 0 for 0 s
w
< s
.
At the s
where (5.92) is satisfied we obtain that:
Ξ
s
w
= 0 and
2
f
w
s
2
w
> 0, (5.104)
so there is a neighborhood V
s
that satisfies:
Ξ
>
f
w
s
w
, if s
w
V
s
/ s
w
< s
Ξ <
f
w
s
w
, if s
w
V
s
/ s
w
> s
.
(5.105)
So for s
V
s
, s > s
the graph of v
CS
lies below the graph of f
w
/s
w
. The point s
minimizes Ξ.
Claim 2: There is another water saturation point s
••
that satisfies (5.92); s
••
maximizes Ξ and satisfies s
••
> s
infl
.
Proof: From Eq.
(5.102) and (5.105) we know that Ξ increases for s
w
> s
. From
Eq.
(5.105), Ξ < fw/s
w
for s
w
V
s
and s
w
> s
,asΞ is continuous we obtain that
s
••
satisfies:
s
••
> s
infl
. (5.106)
If
(5.106) is not satisfied, s
••
is a extremum point for Ξ and satisfies:
Ξ
(
s
••
)
s
w
= 0 and
2
f
w
(
s
••
)
s
2
w
> 0. (5.107)
As Ξ
< f
w
/s
w
for s
w
> s
, and Ξ(s
••
)=f
w
/s
w
, so there is a neighborhood
ˆ
V
••
of
s
••
that satisfies:
Ξ
s
w
>
2
f
w
s
2
w
0 for s
w
ˆ
V
••
/ s
w
< s
••
. (5.108)
Taking the limit s
w
−→ s
••
in Eq. (5.108) we obtain that:
0
=
Ξ
s
w
>
2
f
w
s
2
w
0. (5.109)
For s
••
to satisfy Eq. (5.109), it is necessary s
••
= s
infl
, but from Eq. (5.98) we obtain
that Ξ
(s
infl
) < f
w
(s
infl
)/s
w
,sos
••
> s
infl
.
The point s
••
maximizes Ξ, because Ξ increases for s
w
< s
••
and as Ξ( s
••
)/s
w
= 0
and
2
f
w
(s
••
)/s
2
w
< 0, so there is a neighborhood V
s
••
where Ξ satisfies:
Ξ
<
f
w
s
w
, if s
w
V
s
••
/ s
w
< s
••
Ξ >
f
w
s
w
, if s
w
V
s
••
/ s
w
> s
••
.
(5.110)
130
For s V
s
••
, s > s
••
the graph of Ξ lies above the graph of f
w
/s
w
, so from (5.102) we
obtain that:
Ξ
s
w
> 0, if s
w
V
s
••
/ s
w
< s
••
Ξ
s
w
> 0 < 0, if s
w
V
s
••
/ s
w
> s
••
.
(5.111)
The point s
••
maximizes Ξ.
Claim 3: There is no other point that satisfies
(5.92).
Proof: If this point exists, we denote it by s
•••
,; it should minimize locally Ξ because
Ξ
> f
w
/s
w
for s
w
> s
••
, then from (5.102) we obtain that Ξ decreases. Moreover, Ξ
would satisfy:
Ξ
>
f
w
s
w
, for s
••
< s
w
< s
•••
, (5.112)
and as
2
f
w
/s
2
w
< 0 for all point s
w
> s
infl
(including s
•••
) and as Ξ(s
•••
)/s
w
= 0
there should exist a neighborhood V
s
•••
where
Ξ
>
f
w
s
w
for s
w
V
s
•••
\s
•••
. (5.113)
So from
(5.113), we obtain that Ξ is tangent in s
•••
, moreover from (5.102)
Ξ
s
w
< 0, for s
w
V
s
•••
. (5.114)
Since Ξ
> f
w
/s
w
for s
w
< s
•••
and as Ξ(s
•••
)=f
w
(s
•••
)/s
w
is is necessary that:
Ξ
s
w
<
2
f
w
s
2
w
, for
ˆ
V V
S
•••
(5.115)
where
ˆ
V is a neighborhood where
(5.115) is satisfied.
Taking the limit s
w
−→ s
•••
in (5.115) we obtain
0
=
Ξ( s
•••
)
s
w
2
f
w
s
2
w
< 0. (5.116)
so we conclude that there is no other point that satisfies
(5.92).
There are only two points that satisfy
(5.92) for each fixed T
.
From Prop. 5.6.4, one can prove:
Corollary 5.6.3. For each fixed T in the br, the solution after s
continues as a rar-
efaction in the br.
Proof: From Proposition 5.6.4, s
minimizes Ξ. From eq. (5.96) , we know that
Ξ
(s
w
= 0) < 1, (5.117)
So for s
w
= s
we obtain:
f
w
s
w
= Ξ < 1 <
1 f
w
(s
w
)
1 s
w
=
f
w
(s
w
)
s
w
, (5.118)
131
s
is the smallest value where the rarefaction can be sketched in the br.
Prop. 5.6.4 implies that for each fixed temperature T
there are two saturations
that satisfy Λ
= 0, i.e.
λ
b
s
= v
CS
, but we choose the one that belongs to the CSS and
denote it by s
.
In Fig. 5.9. a, we obtain the water saturation s
, for each T
. We represent this
map of T
into s
, which is used in Item (2) of the following proposition, as:
s
:= s
(T
). (5.119)
Proposition 5.6.5. If T
<
ˆ
T, then
λ
g
T
> v
CS
. Thus there is a shock between (s
w
=
0,T
,u
) to (s
, T
b
,u
+
). Furthermore, the solution continues in the br as a rarefaction.
Proof: The proof consist of two steps:
Figure 5.9: a) Left: The graph of Λ = 0. For each T
, the vertical line crosses the
graph at a saturation s
= s
(T
). b)-Right: The value T
Π
(s
R
) is obtained from the
TCS locus. We plot a horizontal line from s
+
w
, the point where this line intersects the
graph Υ is the point
(T
Π
(s
+
w
),s
+
w
).
(1) The characteristic speed
λ
g
T
is larger than the shock speed v
CS
at (T
L
,s
). From
Figs. 5.9. a and 5.8, we can see that Υ
> 0 at (T
,s
),so
λ
g
T
> v
CS
. Thus there is a
shock between
(0,T
,u
) and (s
, T
b
,u
+
).
(2) The solution continues as a rarefaction in the br. As s
satisfies Λ = 0, this fact
follows from Cor. 5.6.3.
Prop. 5.6.5 yields:
Corollary 5.6.4. When T
tends to T
b
, the saturation s
tends to s
wc
, the connate
water saturation (see Appendix A); thus the solution is continuous when T tends to T
b
between the sr and the br.
132
Proof: Using Λ = 0, where Λ is given by Eq. (5.91), taking the limit of T
L
−→ T
b
,
we obtain:
f
w
(s
w
, T
b
)
s
w
=
f
w
(s
w
, T
b
)
s
w
. (5.120)
The smallest value where
(5.120) is satisfied is s
w
= s
wc
,sos
= s
wc
and the solution
is continuous between the sr and the br.
Remark 5.6.4. When T
= T
b
, the left state lies in the br. This solution was obtained
in [6]. In that case, there was a region with connate water saturation in the br. From
Cor. 5.6.4, our solution agrees with that in [ 6]. Notice that the connate water in this
region is immobile, but this water evaporates when the vaporization shock advances.
Fix s
+
w
< s
. Using the TCS locus, we obtain T
Π
(s
+
w
) as a function of s
+
w
. For each
s
+
w
we draw a horizontal line. We project the intersection of this horizontal line and
the TCS onto the horizontal axis to obtain T
Π
; we denote this mapping by:
T
Π
:= T
Π
(s
+
w
). (5.121)
It is important that T
Π
(s
+
w
) is monotone for s
+
w
< s
.
Proposition 5.6.6. For s
+
w
< s
and T
>
ˆ
T, there is a rarefaction from
(0,T
,u
) to
(0,T
Π
,u
Π
).At(0,T
Π
,u
Π
) the following speeds coincide:
λ
g
T
(T
Π
,u
Π
)=v
CS
(T
Π
,u
Π
;s
+
w
), (5.122)
so there is a left characteristic shock between
(0,T
Π
,u
Π
) and (s
+
w
, T
b
,u
+
) with speed
v
CS
.
Proof: In Fig. 5.9.b, we plot an example of s
+
w
and its respective T
Π
(s
+
w
). Since the
temperature decreases from left to right along the thermal rarefaction wave, from
Rem.
(5.4.3), this wave is a rarefaction.
Corollary 5.6.5. When s
+
w
tends to 0 in br, the temperature T
Π
converges to T
b
; thus
the solution is continuous between the br and the sr.
Proof: Taking the limit of s
+
w
−→ 0 in Eq. (5.89) we obtain:
Υ
(s
w
= 0)=
ρ
g
c
g
ˆ
C
r
+
ρ
g
c
g
ρ
b
g
h
g
h
b
g
ρ
b
g
h
g
h
b
g
+
ˆ
C
r
. (5.123)
We multiply the second term on the right hand side by
(T
T
b
)/(T
T
b
) and we
obtain:
Υ
(s
w
= 0)=
ρ
g
c
g
ˆ
C
r
+
ρ
g
c
g
ρ
b
g
h
g
h
b
g
/
T
T
b
ρ
b
g
h
g
h
b
g
/
T
T
b
+
ˆ
C
r
. (5.124)
133
Taking the limit T
−→ T
b
in (5.124), we obtain:
lim
T
−→T
b
Υ(S
w
= 0)=
ρ
b
g
c
b
g
ˆ
C
r
+
ρ
b
g
c
b
g
lim
T
−→T
b
ρ
b
g
h
g
h
b
g
/
T
T
b
ρ
b
g
h
g
h
b
g
/
T
T
b
+
ˆ
C
r
=
=
ρ
b
g
c
b
g
ˆ
C
r
+
ρ
b
g
c
b
g
ρ
b
g
c
b
g
ˆ
C
r
+
ρ
b
g
c
b
g
= 0. (5.125)
Thus T
= T
b
belongs to the TCS locus. Moreover, Eq. (5.123) defines a curve with
S
R
= 0 and T
in the sr; we can prove that this curve is monotonic (see Figure 5.10),
therefore this is the only solution.
Figure 5.10: Left: plot of Eq. (5.123) for T [373.15K,380K]. Right: the same plot for
T
[373.15K,490K] .
In Eq. (5.121), we find s
= s
(T
); also T
Π
= T
Π
(s
+
w
) from Eq. (5.119), see Fig.
5.9.b. Using s
and T
Π
, four possible solution candidates arise:
(i) T
< T
Π
, s
< s
w
. A shock from (0,T
,u
) to (s
, T
b
,u
+
), continuing to
(s
+
w
, T
b
,u
+
) through a Buckley-Leverett rarefaction.
(ii ) T
< T
Π
, s
> s
+
w
. A shock from (0,T
,u
) to (s
+
w
, T
b
,u
+
) with speed v
CS
.
(iii ) T
> T
Π
, s
< s
+
w
. A rarefaction from (0,T
,u
) to (0,T
Π
,u
Π
) followed by a
shock to
(s
, T
b
,u
+
) continuing to (s
+
w
, T
b
,u
+
) through a Buckley-Leverett rarefaction.
(iv) T
> T
Π
, s
> s
+
w
. A rarefaction from (0, T
,u
) to (0,T
Π
,u
Π
), followed by a
shock to
(s
, T
b
,u
+
) with speed v
CS
.
Proposition 5.6.7. For left temperature T
and fixed right saturation s
+
w
satisfying
T
<
ˆ
T and s
+
w
< s
, (i) and (iii ) do not occur.
134
Proof: From Fig. 5.11. a, we separate the water saturation in the br (the right side
states of the Riemann problem) in two intervals, s
RegI
and s
RegII
. The first one is s
+
w
[s
wc
,s
,]; the second one is s
+
w
[s
w
= 0,s
wc
]. Recall Eq. (5.121), which defines T
Π
as
function of s
+
w
. So we define T
RegI
= T
Π
(s
RegI
); similarly, we define T
RegII
= T
Π
(s
RegII
).
One can verify from Fig. 5.11.a that T
Π
(s
+
w
) is monotone increasing.
Figure 5.11: a)-Left: The map T
Π
defines T
RegI
and T
RegII
. The saturation lies in the
br, which is subdivided in s
RegI
=[s
wc
,s
] and in s
RegII
=[0,s
wc
]. The corresponding
intervals in the sr are T
RegI
and T
RegII
. b)-Right: Mapping from T
RegI
to br. Each
T
T
RegI
defines a s
in the br (see Prop. 5.6.6); notice that s
satisfies s
s
, which
is the saturation at SHB.
Fig. 5.11.b represents the mapping from T
RegI
to the br. Each value of T
T
RegI
defines a value for s
in the br through the function s
(T
) in Eq. (5.119). The function
s
(T
) is not monotone and all values of s
= s
(T
) for T
T
RegI
are larger or
equal to s
, the saturation at the SHB point. Notice also that the mapping s
(T
) for
T
T
RegI
T
RegII
is onto s
RegI
. So we obtain that s
+
w
< s
(T
) for T
T
RegI
,so(i)
and (iii ) do not occur.
Solution: (summarized in Fig 5.13).
(I) For T
L
>
ˆ
T and s
R
> s
infl
> s
. The waves R
T
CS R
s
S
s
with sequence:
L
=(0,T
L
,u
L
)
R
T
(0,
ˆ
T,
ˆ
u)
CS
(s
, T
b
,u
R
)
R
s
(s
§
, T
b
,u
R
)
S
s
(s
R
, T
b
,u
R
)=R, (5.126)
where
(
ˆ
T,s
) is the SHB point and s
§
is given by Eq. (5.83).
(II) For T
L
>
ˆ
T and s
infl
> s
R
> s
. The waves R
T
CS R
s
with sequence:
L
=(0,T
L
,u
L
)
R
T
(0,
ˆ
T,
ˆ
u)
CS
(s
, T
b
,u
R
)
R
s
(s
R
, T
b
,u
R
)=R. (5.127)
(III) For T
Π
< T
L
and s
R
< s
(T
L
). The solution is sketched in Fig. 5.12.a, based
on Props. 5.6.7 and 5.6.6; the waves are R
T
CS with sequence:
L
=(0,T
L
,u
L
)
R
T
(0,T
Π
<
ˆ
T,u
Π
)
CS
(s
R
, T
b
,u
R
)=R. (5.128)
135
Figure 5.12: a)-Left: Riemann Solution. the point T
L
in the horizontal axis represents
(s
L
= 0,T
L
). The rarefaction from (s
L
= 0,T
L
) to (s
L
= 0, T
Π
) is represented by a line
and an arrow to indicate the direction of increasing speed; the rarefaction is followed
by a shock from
(s
w
= 0,T
Π
) to (s
R
, T
b
) with speed v
CS
with construction shown by
dotted lines. b)-Right: the dotted line represents the shock from
(s
L
= 0,T
L
,u
L
) to
(s
R
, T
b
,u
R
) with speed v
CS
. We draw the solution for a left state (s
L
= 0,T
L
), which
we represent by T
L
.
136
(IV) For T
Π
> T
L
and s
R
< s
(T
L
), the solution is sketched in Fig. 5.12.b. It is the
wave CS with sequence:
L
=(0,T
L
,u
L
)
CS
(s
R
, T
b
,u
R
)=R. (5.129)
(V) For T
L
<
ˆ
T and s
infl
> s
R
> s
(T
L
), where the mapping s
= s
(T
L
) is defined
in Eq.
(5.119). We obtain the waves CS R
s
with sequence:
L
=(0,T
L
,u
L
)
CS
(s
, T
b
,u
R
)
R
s
(s
R
, T
b
,u
R
)=R. (5.130)
(VI) For T
L
<
ˆ
T and s
R
> s
infl
> s
. See (5.130). The waves CS R
s
S
s
with
sequence:
L
=(0,T
L
,u
L
)
CS
(s
, T
b
,u
R
)
R
s
(s
§
, T
b
,u
R
)
S
s
(s
R
, T
b
,u
R
)=R, (5.131)
where s
§
is given by Eq. (5.83) with s
+
w
= s
R
.
Remark 5.6.5. We remark that the CS is a double sonic transitional wave, see
[58].
Wave behavior on bifurcations
Notice that the Buckley-Leverett shock S
S
disappears on the line that separates I
from II; we call this line R
S
S
S
bifurcation”. On the line that separates I from VI,
the thermal rarefaction R
T
in sr disappears; we call this line the R
T
S
CS
bifurca-
tion”. On the line that separates II and V, the thermal rarefaction R
T
in the br also
disappears, so this line is the R
T
S
CS
bifurcation”. On the line that separates V
from VI the Buckley-Leverett shock S
S
disappears, so this line is the R
S
S
S
bifur-
cation”. The R
T
S
CS
bifurcation” is the vertical line from SHB separating I, II from
V, VI and the R
T
S
CS
bifurcation” is the horizontal line from S
infl
separating I, VI
from II, V.
On the line that separates II from III the Buckley-Leverett rarefaction R
S
disap-
pears; we call this line the R
S
S
CS
bifurcation”. On the curve that separates III
from IV, the thermal rarefaction R
T
disappears; we call it the“R
T
S
SC
bifurcation”.
Finally, on the curve that separates IV from V the Buckley-Leverett rarefaction R
S
disappears; we call it the S
SC S
R
S
bifurcation”.
5.6.4 Riemann Problem D
We inject pure water at temperature T
L
< T
b
, i.e, the left state is (0,T
L
,u
L
); on the
right we have pure steam at temperature T
R
> T
b
.
Before describing our proposed Riemann solution, it is necessary to prove that
there is no possible Riemann solution with a direct shock between sr and the wr.
Using the RH condition
(5.36)-( 5.37) with s
+
w
= 0 and T
b
replaced by T
+
> T
b
,weob-
tain this hypothetical “complete water evaporation shock”, labelled CWES, with speed
v
CWE S
. The superscript + () in the following equations represents the temperature
137
Figure 5.13: Phase diagram for Riemann Problem C. The dotted line delimits the
physical range, the dashed lines are bifurcation loci. The continuous curves are parts
of the TCS and the CSS bifurcations loci. The horizontal axis represents the left
states
(0,T
L
,u
L
) in sr; the vertical axis represents the right states (s
R
, T
b
,u
R
) in br.
The solutions are given in I-VI, Sec. 5.6.3.
T
+
(T
). The speed of such shock between (s
w
= 1,T
,u
w
) to (s
w
= 0,T
+
> T
b
,u)
would be:
v
CWE S
(T
,u
;s
+
w
= 0,T
+
)=
u
w
ϕ
ρ
w
(h
+
g
h
w
)
H
+
r
H
r
+
ρ
w
(h
+
g
h
w
)
. (5.132)
Proposition 5.6.8. Complete Evaporation. For any T
< T
b
< T
+
, if there exists
a complete water evaporation shock from
(1,T
,u
) to (0,T
+
> T
b
,u
+
) with speed
v
CWE S
(T
,u
;s
+
w
= 0,T
+
) given by (5.132), then this shock satisfies:
v
CWE S
> v
w
T
, (5.133)
where v
w
T
is the speed of thermal discontinuity given by Eq. (5.32).
Proof: The speed in the liquid water region v
w
T
is written as:
v
w
T
=
u
w
ϕ
C
w
ˆ
C
r
+ C
w
=
u
w
ϕ
C
w
ˆ
C
r
+ C
w
(
T
+
T
)
(
T
+
T
)
=
u
w
ϕ
ρ
w
(h
+
w
h
w
)
H
+
r
H
r
+
ρ
w
(h
+
w
h
w
)
.
Using v
CWE S
given by (5.132) and the relationship
ρ
w
(h
+
g
h
w
) >
ρ
w
(h
+
w
h
w
),it
follows that v
CWE S
> v
w
T
.
If this shock exists, it would not satisfy the Oleinik condition for entropy [53].
Therefore we conclude that instead of a shock there is a br between the
() and (+)
state. The solution is constructed using results from Sections 5.6.1 and 5.6.2. Since
s
+
w
= 0, from Sec. 5.6.2 we obtain:
138
Proposition 5.6.9. The saturation s
††
in (5.80) is larger than
ˆ
s (defined in (5.62)),
thus there is a rarefaction from
(s
††
, T
b
,u
+
) to (
ˆ
s, T
b
,u
+
).
Proof: For s
w
= 1, we obtain from Eq. (5.39) and (5.58) that v
WES
and v
VS
satisfy:
v
WES
=
u
b
ϕ
ρ
w
(h
b
w
h
w
)
ˆ
H
b
r
ˆ
H
r
+
ρ
w
(h
b
w
h
w
)
=
u
b
ϕ
C
w
ˆ
C
r
+ C
w
,
v
VS
=
u
b
ϕ
ρ
w
(h
+
g
h
b
w
)
ˆ
H
+
r
ˆ
H
b
r
+
ρ
w
(h
+
g
h
b
w
)
;
we can show as in Proposition 5.6.8 for s
w
= 1 that
v
VS
> v
WES
. (5.134)
The inequality Eq.
(5.134) is satisfied for s
w
, since that f
w
s
w
.
We know that s
††
is given by (5.80) and for this saturation f
w
(s
††
) > s
††
, so Eq.
(5.134) is satisfied from s
w
= 1 to at least s
w
= s
††
.Asf
w
/s
w
increases for s
w
decreasing from s
w
= 1 to s
infl
, the water saturation s
††
that satisfies (5.80) is larger
than
ˆ
s, where
ˆ
s satisfies
(5.62), because f
w
s
w
and (5.134) is satisfied.
Corollary 5.6.6. The solution in the br consists of a rarefaction from (s
††
, T
b
,u
+
) to
(
ˆ
s, T
b
,u
+
).
Proof: As s
††
>
ˆ
s and from Section 5.6.1, we know that the solution consists of a
rarefaction from s
w
= 1 to s
w
=
ˆ
s in the br, we conclude that exists a rarefaction from
s
††
to
ˆ
s in this region.
Solution. The solution consists of the waves WES R
s
CS with sequence:
L
=(1,T
L
,u
L
)
WES
−−→ (s
††
, T
b
,u
R
)
R
s
(
ˆ
s, T
b
,u
R
)
SC
(0,T
R
,u
R
)=R. (5.135)
5.6.5 Riemann Problem E
We inject pure steam at temperature T
L
> T
b
, i.e, L =(0,T
L
> T
b
,u
L
); on right we
have pure water at T
R
< T
b
, i.e., the right state is R =(1,T
R
< T
b
,u
R
). We use results
from Section 5.6.4 to obtain the solution.
We remark that the vaporization shock, VS, is the reverse of the condensation
shock, CS, so there could exist a hypothetical “complete condensation shock”, labelled
CC S, with speed v
CC S
. This shock would be obtained using the RH condition (5.43)-
(5.44) with the () state replaced by (sw = 1,T
+
< T
b
,u
+
), and the (+) state replaced
by
(s
w
= 0,T
> T
b
,u
). The speed of such shock would be:
v
CC S
(T
,u
;s
+
w
= 1,T
+
)=
u
ϕ
ρ
g
(h
+
g
h
w
)
H
+
r
H
r
+
ρ
+
g
(h
+
g
h
w
)
. (5.136)
The following fact indicates that instead of this shock, there is always a br between
the sr and the wr.
139
Proposition 5.6.10. Complete Condensation. For any T
T
b
the complete conden-
sation shock from
(s
w
= 0, T
,u
) to (s
w
= 1, T
+
,u
+
) with speed v
CC S
(T,u
;s
+
w
= 1)
satisfies:
v
CC S
>
λ
g
T
(T
,u
). (5.137)
Proof: We divide the numerator and denominator of
(5.136) by (T
+
T
):
v
CC S
(s
w
= 1)=
u
ϕ
ρ
+
g
(h
+
g
h
w
)
/
(
T
+
T
)
ˆ
C
r
+
ρ
+
g
(h
+
g
h
w
)
/
(
T
+
T
)
>
u
ϕ
ρ
+
g
(h
+
g
h
g
)
/
(
T
+
T
)
ˆ
C
r
+
ρ
+
g
(h
+
g
h
g
)
(
T
+
T
)
>
u
ϕ
ρ
+
g
c
+
g
ˆ
C
r
+
ρ
+
g
ρ
+
g
=
λ
g
T
, (5.138)
the inequality above are valid because h
0
g
> h
0
w
and c
g
=
2
h
g
/T
2
< 0 for T
> T
b
.
Corollary 5.6.7. There is a br between the states (s
w
= 0,T
,u
) and (s
w
= 1,T
0
,u
w
).
Proof: From Proposition 5.6.10, there exists a br between
(s
w
= 0,T
,u
) and
(s
w
= 1,T
0
,u
w
), because if this region did not exist, as Eq. (5.137) is satisfied it would
be possible to sketch a rarefaction from
(s
w
= 0,T
,u
) to (s
w
= 0,T
b
,u
b
) in the br.
In [6], Bruining et al. obtained the solution for injection of water and steam at
boiling temperature in a porous rock with water at temperature T
0
. In that work, the
water and rock enthalpies were made to vanish at temperature T
0
, so the formulae
for shock speeds are slightly different from formulae in present paper; however both
are equivalent since the enthalpy is defined in up to a constant. In that paper, two
saturations denoted by s
and s
††
were found both satisfying Eq. (5.80); the choice for
s
††
was also made. In [6], cases I and II are correct, but there are some mistakes that
influence the solution in case III. The first relevant mistake is the statement that the
saturation S
maximizes v
SC F
(Remark 11). The correct statement is:
Proposition 5.6.11. There are two saturation values that satisfy Eq. (3.7) in
[6 ] for
each fixed T
< T
b
. The smallest S
minimizes v
SC F
. The other S
††
maximizes v
SC F
, but
it is irrelevant.
Moreover the solution is stable if we vary the left state, it follows that:
Corollary 5.6.8. In the limit as S
L
tends to zero, the wave speed v
b
g,w
in the br given
by Eq. (2.8) in
[1 ] converges to zero, so the solution for the Riemann problem reduces a
cooling discontinuity in the liquid water region.
From Proposition 5.6.11 and Corollary 5.6.8, we see that the Figures 3.1 and 3.4 in
[6 ] contain an error; we correct them in Figure 5.14.
140
Figure 5.14: a)-Left: Corrected schematic bifurcation near S
(for fixed u
0
, T
0
versus
S
b
w
. The shock speed SCF is minimum at S
; the shock speed v
b
g,w
tends to zero when
S
b
w
tends to 1. b)-Right: The corrected structure of steam-water zone below solid
curve marked by v
b
s
, v
SC F
T
, v
SC F
given in Eqs (2.7), (2.24) with S
w
= S
, Eq. (2.24) with
S
< S
inj
< S
, and Eq. (2.8) respectively. The figures are not drawn to scale.
The wave speed diagram in Figure 4.5 of
[6 ] contains an error also. The correct
diagrams are given in Figure 5.15. There are two diagrams because the saturation
shock speed and the thermal shock are different for injected saturations larger than
S
. These diagrams are drawn out of scale for illustrative purposes, because the
characteristic speed v
g
s
can be much larger than the other wave speeds. Figs. 3.1 and
3.4 in
[6 ] contain an error also; those figures were summarized in Figure 4.5 in [ 6],
which is corrected here in Figure 5.15. The Riemann solution for the cases I and II
are correct, however the Riemann solution for case III contains an error. The mistake
is the statement that the saturation shock speed
λ
b
s
is faster than v
SC F
in cases I and
II. The correct statement is that v
b
g,w
speed converges to zero if the water saturation
at the left state tends to zero, as summarized in Corollary 5.6.8 above. The thermal
wave speed is drawn in Figure 5.15.a
) and in 5.15.b) we draw the saturation wave
speed.
141
Figure 5.15: a)-Left: The diagram represents the thermal wave speed; b)-Right: The
diagram represents the saturation wave speed. They correct the diagram presented
in Figure 4.5 of
[6 ]. In both, the solid curve represents the wave speeds used in the
Riemann solution. The characteristic speed v
b
s
, Eq. (2.7), is the Buckley-Leverett
characteristic speed in the br; the shock speed v
b
g,w
is the HISW speed, Eq. (2.8),
and represents the Buckley-Leverett shock for a water saturation s
w
to s
w
= 0; v
SC F
is the condensation shock speed between the br and the wr, Eq. (2.24); v
b,0
w
is the
cooling contact discontinuity speed in the liquid water region, Eq.
(2.15). Notice that
the temperature shock speed tends to v
W
T
when water saturation tends to S
w
= 1
(liquid water region) and the saturation shock speed tends to zero, thus the solution
converges to the solution in the liquid water region.
142
The solution diagram for case III given in Figure 4.4 of [6] must be modified. The
strength of saturation shock tends to zero when the injection saturation tends to 1,
while the speed of cooling discontinuity does not change. In Figure 5.16 we show the
schematic solution for case III for three different injection saturations.
Figure 5.16: Steam injection for three high water saturation values. We get a constant
state upstream, the Buckley-Leverett saturation shock and the cooling discontinuity.
These waves have distinct speeds; notice that v
b
g,w
tends to zero when the injection
saturation tends to 1, while the speed of cooling discontinuity does not change.
To complete the Riemann solution, we obtain a relationship between s
and s
:
Proposition 5.6.12. The water saturation s
, defined in (5.119), obtained in the br by
a shock from
(0,T
,u
) to (s
, T
b
,u
+
) satisfies the following inequality:
s
< s
,
so from
(s
, T
b
,u
+
) the solution continues as a rarefaction to (s
, T
b
,u
+
).
Proof: We analyze the relationship between the shock speeds v
CS
and v
SC F
.It
is possible show that v
SC F
is faster than v
CS
. Since (T
,s
) belongs to CSS so v
CS
satisfies:
v
CS
=
u
b
ϕ
f
w
(s
, T
b
)
s
w
,
143
and v
SC F
satisfies Eq. (5.80), we obtain that:
f
w
(s
, T
b
)
s
w
<
f
w
(s
, T
b
)
s
w
,
as f
w
/s
w
is monotone increasing for s
wc
< s
w
< s
infl
w
, we obtain the result.
Solution. We obtain the Riemann solution using the results in [6], [39] and
(5.131), (5.126).
For T
L
<
ˆ
T. As in
(5.131), there is a shock from L =(0,T
,u
) to (s
, T
b
,u
+
)
where s
= s
(T
L
) is given by (5.119) and s
satisfies Eq. (5.80). The waves are
CS
R
s
SCF with sequence:
L
=(0,T
L
,u
L
)
CS
(s
, T
b
,u
R
)
R
s
(s
, T
b
,u
R
)
SC F
(s
R
, T
b
,u
R
)=R. (5.139)
For T
L
>
ˆ
T. The waves R
T
CS R
s
SCF with sequence:
L
=(0,T
L
,u)
R
T
(0,
ˆ
T,u)
CS
(s
, T
b
,u
b
)
R
s
(s
, T
b
,u
b
)
SC F
(s, T < T
b
,u
R
)=R,
(5.140)
where
(
ˆ
T,s
) is the SHB point.
144
CHAPTER 6
Riemann solution for problem of Chapter 4 for V
L
in III
and IV.
Here we utilize results of Chapter 4 and Chapter 5. First we can identify subre-
gions
L
5
and L
6
in III
IV. The only way to reach spl region from III and IV is by a
Buckley-Leverret shock, see Section 4.9.3. Notice that the boundary between III and
IV is the coincidence curve
λ
s
=
λ
e
, in which
λ
s
and
λ
e
are given by Eqs. (4.115) and
(4.117). Moreover, there exists a coincidence curve between the evaporation wave and
the hot isothermal steam water, HISW, with speed v
b
g,w
and given by Eq. (5.28);we
denote this coincidence curve by
C
BLE
, see Fig. 6.1.
In next Riemann solutions, s represents the water saturation.
6.1 V
L
in L
6
In this region, HISW is slower than the evaporation wave, so we can reach the spl
directly by a HISW. It is easy to prove that:
Lemma 6.1.1. The following inequalities are valid for
(s
w
, T) in L
6
:
v
b
g,w
(s
w
, T) < v
WES
(s
w
= 1,T; s
††
) < v
w
T
, (6.1)
where s
††
is obtained from Eq. (5.80).
The Riemann solution consists of the waves S
BL
WES R
s
CS with se-
quence:
L
=(s
L
, T
L
,u
L
)
BL
(1,T
L
,u
L
)
WES
−−→ (s
††
, T
b
,u
R
)
R
s
(
ˆ
s, T
b
,u
R
)
SC
(0,T
R
,u
R
)=R.
(6.2)
145
6.2 V
L
in L
5
In this region, HISW is faster than the evaporation wave. From a state in
L
5
, we reach
C
BLE
using a condensation rarefaction. From this state in the coincidence curve, we
can reach the spl directly by a HISW.
The Riemann solution consists of the waves R
e
S
BL
WES R
s
CS with
sequence:
L
=(s
L
, T
L
,u
L
)
R
s
(s, T,u)
BL
(1,T,u)
WES
−−→ (s
††
, T
b
,u
R
)
R
s
(
ˆ
s, T
b
,u
R
)
SC
(0,T
R
,u
R
)=R,
(6.3)
where
(s, T) are the states in the coincidence curve C
BLE
obtained from the evapora-
tion rarefaction wave from
(s
L
, T
L
) and u is the respective speed.
L
5
L
6
C
5
BLE
Figure 6.1: Subregions L
5
and L
6
. C
BLE
is the coincidence curve v
b
g,w
=
λ
e
. For states
in
L
5
, we have that v
b
g,w
<
λ
e
; for states in L
6
, we have that v
b
g,w
>
λ
e
.
146
L
R
1
S
BL
R
S
BL
L
R
e
1
2
Figure 6.2: Riemann solutions in phase space, omitting the surface shown in Fig. 4.1.
a) Left: Solution
(6.2) for V
L
∈L
6
, Sec. 6.1. b) Right: Solution (6.3) for V
L
∈L
5
,
Sec. 6.2. The numbers 1 and 2 indicate intermediate states V in wave sequence. The
waves after spl situation are not represented in this figure, but can be obtained in the
previous Chapter.
147
CHAPTER 7
Summary and Conclusions
We have described a global formalism for balance laws of form (2.1). We show a
systematic theory for the Riemann solution for general Riemann problems for a wide
class of balance equations with phase changes.
We have also obtained the solution of the Riemann problem for the injection of a
mixture nitrogen/steam/water into a porous rock filled with steam above boiling tem-
perature. We show a systematic theory for the Riemann solution for a 3
× 3 balance
system. The set of solutions depends L
1
continuously on the Riemann data.
We have also described completely all possible solutions of the Riemann problem
for the injection of a mixture of steam and water in several proportions and tempera-
ture into a core filled with a different mixture of steam and water in all proportions,
(of course, the temperature must be lower than the thermodynamical critical tem-
perature of water). The set of solutions depends L
1
continuously on the Riemann
data. We found several types of shock between regions and systematized a scheme to
find the solution from these shocks. A new type of shock, the evaporation shock, was
identified.
As a future work we would like obtain the Riemann solution for the problem de-
scribed in Chapter 3. We are interested also in considering the gravitational and
capillarity pressure effects and the effects of pressure changes in terms of physical
properties.
148
Bibliography
[1] BAYLISS, A., MATKOWSKY,B.J,AND ALDUSHIN,A.P.,Dynamics of hot spots in
solid fuel combustion, Physica D, Vol. 166, Vol. 1-2, pp. 104-130, 2002.
[2] B
EEK,W.J.AND MUTZALL,M.K.,Transport Phenomena, John Willey and Sons.
[3] B
IRD, R.B., STEWART, W. E., AND LIGHFOOT,E.N.,Transport Phenomena, John
Wiley and Sons (1960).
[4]
BRUINING, J, DIETZ, D.N., EMKE, A., METSELAAR, G., and SCHOLTEN, J.W., Im-
proved recovery of heavy oil by steam with added distillables,3
rd
European Meet-
ing on Improved Oil Recovery, Rome, Proc I, 371-378 (1985).
[5] B
RUINING,J.,DUIJN,C.J.VAN, Uniqueness Conditions in a Hyperbolic Model for
Oil Recovery by Steamdrive, Computational Geosciences, Vol. 4, pp 65-98, 2000.
[6] B
RUINING,J.,MARCHESIN,D.E VAN DUIJN,C.J.,Steam Injection Into Water-
Saturated Porous Rock, Computational and Applied Mathematics, vol. 22, N.3, pp.
359-395, 2003.
[7] B
RUINING,J., Multiphase flow in porous media, unpublished monograph.
[8] B
RUINING,J.,VAN DUIJN,C.J.AND MARCHESIN,D., A Hyperbolic Model For
Steamdrive and Volatilization in Porous Media, 2004, preprint.
[9] B
RUINING,J.AND VAN DUIJN,C.J., Traveling waves in a finite condensation
rate model for steam injection, 2005, preprint.
[10] B
RUINING,J.,MARCHESIN,D.,Nitrogen and Steam Injection in a Porous
Medium with Water, to appear to Transport in Porous Media (2006 ).
[11] B
RUINING,J.,MARCHESIN,D.,Maximal oil recovery by simultaneous conden-
sation of alkane and steam, Submitted to Phys. Rev. E, 2006.
149
[12] BUCKLEY, S.E, AND LEVERETT,M.C.,Mechanism of fluid displacement in
sands,Trans. AIME, 107-116.
[13] C
OATS,K.H.,A highly implicit steamflood model, Soc. Pet. Eng. J., Oct. 1978,
369-383.
[14] C
OATS,K.H.,In-situ combustion model, Soc. Pet. Eng. J., Dec. 1980, 533-554.
[15] C
OLOMBO, R. M., AND CORLI,A.,Continuous dependence in conservation laws
with phase transitions, Siam J. Math. Anal., Vol. 31, No. 1, pp. 34-62, 1999.
[16] C
ROOKSTON, R.B., CULHAM, W. E., AND CHEN,W.H.,A numerical simulation
model for thermal recovery processes, Soc. Pet. Eng. J., Feb. 1979, 37-58; Trans.,
AIME, 267.
[17] C
UMMINGS,J.,Groundwater Remediation at DNAPL sites,
http://www.epa.gov/epaoswer/osw/meeting /pdf02/cummings.pdf
[18] D
AVIS,E.L.,Steam injection for soil and aquifer remediation, EPA/540/S-
97/505, January 1998.
[19] D
AVIS,E.L.,How heat can enhance in-situ soil and aquifer remediation: im-
portant chemical properties and guidance on choosing the appropriate technique,
EPA/540/S-97/502, April 1997.
[20] D
AFERMOS,C.,Hyperbolic Conservation Laws in Continuum Physics, Springer
Verlag, 1 edition, 443 pgs, 2000.
[21] EPA R
EPORT, Pump and Treat: Groundwater Remediation, EPA/625/R-95/005,
July 1996.
[22] F
AYERS F.J., Some theoretical results concerning the displacement of a viscous
oil by a hot fluid in a porous medium, Fluid Mech.,13 (1962).
[23] L
AKE, L.W., POPE, G.A., CAREY,G.F.,AND SEPEHRNOORI,K.,Isothermal,
Multiphase, Multicomponent Fluid-Flow in Permeable Media, Part I: Description
and Mathematical Formulation, In Situ J., 8 (1), 1-40, 1984.
[24] I
SAACSON, E., MARCHESIN,D.,PLOHR,B.,AND PALMEIRA,F.,A Global for-
malism for nonlinear waves in conservation laws , Comm. Math. Phys., vol. 146,
pp. 505-552, 1992.
[25] I
SAACSON,E., MARCHESIN,D.,PLOHR,B.,AND TEMPLE,J.,B.,The Riemann
Problem near a hyperbolic singularity: Classification of quadratic Riemann prob-
lems I, SIAM Journal of Appl. Math., 48, 1009-1032, 1988 .
[26] I
SAACSON, E., MARCHESIN,D.,PLOHR,B.,AND TEMPLE,J.,B.,Multiphase
flow models with singular Riemann problems, Mat. Apl. Comput., vol. 11, n. 2, pp.
147-166, 1992.
150
[27] DA MOTA, J.C., DANTAS,W.B,GOMES M.E.S., AND MARCHESIN,D.,Combus-
tion Fronts in Petroleum Reservoirs, Mat. Cont., v. 8, n.1, pp. 129-140, 1995;
[28]
DA MOTA, J.C., DANTAS,W.B.,AND MARCHESIN,D.,Travelling waves for
combustion in porous media, Int. Series of Num. Math., Birkhauser Verlag, v.
129, n.2, pp. 177-187, 1999.
[29]
DA MOTA, J.C., SOUZA, A.J., AND MARCHESIN,D.,Riemann solution for a
model of combustion in two-phase flow in porous media, Int. Series of Num. Math.,
v.141, n.2, pp. 683-692, 2001.
[30]
DA MOTA,J.C.,SOUZA,A.J.,GARCIA, R. A., AND TEIXEIRA,P.W.,Oxidation
Fronts in a Simplified Model for Two-Phase Flow in Porous Media, Mat. Cont.,
Rio de Janeiro, v. 22, n. I, pp. 67-82, 2002.
[31]
DA MOTA,J.C.,DANTAS,W.B.AND MARCHESIN,D.,Combustion Fronts in
Porous Media, SIAM Journal on Appl. Math., USA, v. 62, n. issue 6, pp. 2175-
2198, 2002.
[32]
DA MOTA,J.C.AND SCHECTER,S.,Combustion Fronts in Porous Media with
Two Layers, Journal of Dyn. and Diff. Equat., USA, accepted 2006.
[33] F
IROZABADI,A., Thermodynamics of Hydrocarbon Reservoirs, Macgraw-Hill,
1999.
[34] G
ODLEWSKI, E., AND RAVIART, P.A., Numerical Approximation of Hyperbolic
Systems of Conservation Laws , Springer, 1 edition, 1997.
[35] K
ONDRASHOV,A.V.,Properties of certain phase transition fronts in geothermal
porous reservoirs, Fluid Dynamics, Vol. 39, N. 6 pp. 961-971, 2004.
[36] K
RONER,D.,Numerical Schemes for Conservation Laws, John Wiley and Sons,
1997.
[37] K
ULIKOVSKII,A.G.,Evaporation and Condensation Front in Porous Media,
Fluid Dynamics, Vol. 37, N. 5, pp. 740-746, 2002.
[38] K
ULIKOVSKII,A.G.,Phase Transitions in Flow through a Heat-Conducting
Porous Skeleton, Fluid Dynamics, Vol. 39, N. 3 pp. 429-434, 2003.
[39] L
AMBERT,W.,BRUINING,J.AND MARCHESIN,D.,Erratum: Steam Injection
Into Water-Saturated Porous Rock, to appear in Computational and Applied Math-
ematics, 2006.
[40] L
AMBERT,W.,MARCHESIN,D.,BRUINING,J.AND ALBUQUERQUE,D.,Numer-
ical method for nitrogen and steam injection into porous medium with water, Rio
Oil Gas 2004, available in CD-ROM.
151
[41] LAMBERT,W.,MARCHESIN,D.AND BRUINING,J.,The Riemann Solution of
the Balance Equations for Steam and Water Flow in a Porous Medium, preprint
IMPA-A426, available at http://www.preprint.impa.br, pp. 1
41, (2006).
[42] L
AMBERT,W.,MARCHESIN,D.AND BRUINING,J.,Numerical simulation of
steam and nitrogen injection into porous medium, 2004, preprint
[43] L
AX,P.,Hyperbolic systems of conservation laws, II, Comm. Pure and Appl.
Math., 10 (1954), 537-566.
[44] L
EE,F.J.AND SEARS,F.W.,Termodinˆamica, Ao Livro T
´
ecnico S. A. e Editora
da Universidade de S
˜
ao Paulo, Rio de Janeiro, 1969, in portuguese.
[45] L
EFLOCH,P.G.,Hyperbolic Systems of Conservation Laws, Lectures Notes in
Mathematics, Birkhauser, 2002.
[46] L
EVEQUE,R.J.,Numerical Methods for Conservation Laws, Birkhauser, 2 edi-
tion, 1994.
[47] L
IU,T.P.,The Riemann problem for general 2 X 2 conservation Laws, Transac-
tions of A.M.S., 199 (1974), 89-112.
[48] L
IU,T.P.,The Riemann Problem for general systems of conservation laws, J. Diff.
Equations, 18 (1975), 218-234.
[49] M
ACKAY, D.M. AND J.A. CHERRY, Groundwater Contamination: Limits of
Pump-and-Treat Remediation, Environ. Sci. Technol. 23(6): 630-636, 1989.
[50] M
ARX,J.W.AND LANGEHEIM, R.H., Reservoir heating by hot fluid injection,
Trans. AIME, 1959, 216, 312-315.
[51] N
EWMARK, R. L., AND AINES,R.D.,They all like it hot: faster cleanup of con-
taminated soil and groundwater, http://www.llnl.gov/str/Newmark.html.
[52] N
EWMARK, R. L., AND AINES,R.D.,Dumping pump and treat: rapid cleanups
using thermal technology, AICHE 1997 Spring Meeting, Houston, TX, March 10-
12, 1997.
[53] O
LEINIK,O.,Discontinuous Solutions of Nonlinear Differential Equations, Usp.
Mat. Nauk. (N. S.), Vol. 12, pp 3-73, 1957. English transl. in Amer. Math. Soc.
Transl. Ser. 2, 26, 95-172.
[54] P
OPE, G.A., CAREY,G.F.,AND SEPEHRNOORI,K.,Isothermal, Multiphase,
Multicomponent Fluid-Flow in Permeable Media, Part II: Numerical Techniques
and Solution, In Situ J., 8 (1), 41-97, 1984.
[55] P
RATS,M.,Thermal Recovery, Society of Petroleum Engineers, Henry L. Do-
herty Series, Monograph Vol. 7, 1986.
152
[56] RICCI, J. E., The phase rule and heterogeneous equilibrium, Dover Publications,
INC., New York, 1966.
[57] R
ICHARDSON,C.N.AND TSYPKIN,G.G.,Nonlinear effects in water-vapor ex-
traction from a geothermal reservoir,Journal of Engineering Physics and Thermo-
physics, Vol. 77, No. 2, 2004.
[58] S
CHECTER,S.,PLOHR,B.AND MARCHESIN,D., Classification of Codimension-
One Riemann Solutions , Journal of Dynamics and Differential Equations, Vol. 13,
No. 3. pp. 523-588, 2001.
[59] S
ERRE,D.,Systems of Conservation Laws, 2 volums, Cambridge University
Press, 1999.
[60] S
MOLLER,J.,Shock Waves and Reaction-Diffusion Equations, Springer-Verlag,
1983.
[61] S
OUZA,A.J.,Stability of Singular Fundamental Solutions Under Perturbations,
Matem. Aplic. Comput., 1992, 11, 73-115.
[62] S
OUZA,A.J.,AKKUTLU,I.Y.,AND MARCHESIN,D,Wave Sequences for Solid
Fuel In-Situ Combustion in Porous Media, Computational and Applied Mathe-
matics accepted in 2006.
[63] T
SYPKIN, G.G., AND BREVDO, L., , A phenomenological model of the increase
in solute concentration in ground water due to evaporation, Transport in Porous
Media, vol. 37, pp. 129-151, 1999.
[64] T
SYPKIN,G.G.,Existence of the frontal regime of water-steam phase transition
in hydrothermal reservoir, Fluid Dynamics, vol. 35, No. 6, pp. 901-909, 2000.
[65] T
SYPKIN,G.G.AND CALORE,C.,Role of capillary forces in vapour extraction
from low-permeability, water saturated geothermal reservoirs, Geothermics, No
32, pp. 219-237, 2003.
[66] T
SYPKIN,G.G.AND WOODS,A.W.,Vapour extraction from a water-saturated
geothermal reservoir, J. Fluid Mech., vol. 506, pp. 315-330, Cambridge University
Press, 2004.
[67] T
SYPKIN,G.G.AND WOODS,A.W.,Precipitate formation in a porous rock
through evaporation of saline water, J. Fluid Mech., vol. 537, pp. 35-53, Cam-
bridge University Press, 2005.
[68] W
YLEN,G.J.,Thermodynamics, John Wiley and Sons, 1959.
[69] W
ENDROFF,B.,The Riemann problems for materials with non-convex equations
of state I., J. Math. Anal. Appl. 38, 454-466, (1972).
[70] W
ENDROFF,B.,The Riemann problems for materials with non-convex equations
of state II., J. Math. Anal. Appl. 38, 454-466, (1972).
153
154
APPENDIX A
Physical quantities; symbols and values for the Nitrogen
Problem
Table 2, Summary of physical input parameters and variables
Physical quantity Symbol Value Unit
Water, steam fractional functions f
w
, f
g
Eq. (3.15) . [m
3
/m
3
]
Porous rock permeability k 1.0 × 10
12
. [m
3
]
Water, steam relative permeabilities k
rw
, k
rg
Eq. (A.10). [m
3
/m
3
]
Water, steam end point permeabilities k
rw
, k
rg
Eq. (A.10. a), (A.10.b). [m
3
]
Pressure p
at
1.0135×10
5
. [Pa]
Water Saturation Pressure p
sat
Eq. (A.7). [Pa]
Water, steam phase velocity u
w
, u
g
Eq. (3.14) . [m
3
/(m
2
s)]
Total Darcy velocity u u
w
+ u
g
, Eq. (3.16) . [m
3
/(m
2
s)]
Effective rock heat capacity C
r
2.029×10
6
. [J/(m
3
K)]
Steam and nitrogen enthalpies h
gW
, h
gN
Eqs. (A.3), (A.4). [J/m
3
]
Water enthalpy h
W
Eq. (A.2). [J/m
3
]
Rock enthalpy H
r
Eq. (A.1). [J/m
3
]
Water, steam saturations s
w
, s
g
Dependent variables. [m
3
/m
3
]
Connate water saturation s
wc
0.15. [m
3
/m
3
]
Temperature T Dependent variable. [K]
Water, steam thermal conductivity
κ
w
,
κ
g
0.652 , 0.0208. [W/(mK)]
Rock, composite thermal conductivity
κ
r
,
κ
1.83 , Eq. (4.9) . [W/(mK)]
Water, steam viscosity
µ
w
,
µ
g
Eq. (A.5) , Eq. (A.6) . [Pa s]
Steam and nitrogen densities
ρ
gw
,
ρ
gn
Eq. (A.8. a), (A.8.b). [kg/m
3
]
Constant water density
ρ
W
998.2. [kg/m
3
]
Steam and nitrogen gas composition
ψ
gw
,
ψ
gn
Dependent variables. []
Universal gas constant R 8.31 [J/mol /K ]
Nitrogen and water molar masses M
N
, M
W
0.28, 0.18 [kg/mol ]
Rock porosity
ϕ
0.38. [m
3
/m
3
]
155
A.1 Temperature dependent properties of steam and
water
The rock enthalpy H
r
and
ˆ
H
r
are given by:
H
r
=(1
ϕ
)C
r
(T
¯
T
) and
ˆ
H
r
= H
r
/
ϕ
. (A.1)
The water enthalpy per mass unit h
W
is:
h
W
= C
W
T/
ρ
w
. (A.2)
The steam enthalpy h
gW
[J/kg] as a function of temperature is approximated by
h
gW
(
T
)
=
2.20269×10
7
+ 3.65317 × 10
5
T 2.25837×10
3
T
2
+ 7.3742 T
3
1.33437 × 10
2
T
4
+ 1.26913 × 10
5
T
5
4.9688 × 10
9
T
6
h
w
. (A.3)
The nitrogen enthalpy h
gN
[J/kg] as a function of temperature is approximated by
h
gN
(
T
)
=
975.0T + 0.0935 T
2
0.476 × 10
7
T
3
h
gN
. (A.4)
These constants enthalpies
h
w
and h
gN
are chosen such that h
w
(
T
)
, h
gN
(
T
)
vanish at
a reference temperature
T = 293K.
The temperature dependent liquid water viscosity
µ
w
[Pas] is approximated by
µ
w
= 0.0123274+
27.1038
T
23527.5
T
2
+
1.01425×10
7
T
3
2.17342×10
9
T
4
+
1.86935×10
11
T
5
.
(A.5)
We assume that that the viscosity of the gas is independent of the composition.
µ
g
= 1.8264×10
5
T
T
b
0.6
. (A.6)
The water saturation pressure as a function of temperature is given as
p
sat
= 10
3
(175.776 + 2.29272T 0.0113953 T
2
+ 0.000026278 T
3
0.0000000273726T
4
+ 1.13816 × 10
11
T
5
)
2
(A.7)
The corresponding concentrations
ρ
gw
,
ρ
gn
are calculated with the ideal gas law:
ρ
gw
=
M
W
p
sat
RT
,
ρ
gn
=
M
N
p
at
p
sat
RT
, (A.8)
The pure phase densities are:
ρ
gW
(T)=
M
W
p
at
RT
,
ρ
gN
(T)=
M
N
p
at
RT
, (A.9)
where the gas constant R
= 8.31[J/mol /K], M
W
and M
N
are the the nitrogen and
water molar mass.
For simplicity the liquid water density is assumed to be constant at 998.2.kg
/m
3
.
156
A.2 Constitutive relations
The relative permeability functions k
rw
and k
rg
are considered to be power functions
of their respective saturations, i.e.
k
rw
=
k
rw
s
w
s
wc
1s
wc
s
gr
n
w
0
,k
rg
k
rg
=
s
g
s
gr
1s
wc
s
gr
n
g
1
for s
wc
s
w
1,
for s
w
< s
wc
.
(A.10)
For the computations we take n
w
= 2 and n
g
= 2. The end point permeabilities k
rw
,k
rg
are 0.5 and 0.95 respectively. The connate water saturation s
wc
is given in the table.
157
APPENDIX B
Physical quantities; symbols and values for the steam
injection
B.1 Temperature dependent properties of steam and
water
We use reference [1] to obtain all the temperature dependent properties below. The
water and steam densities used to obtain the enthalpies are defined at the bottom.
The steam enthalpy h
g
[J/kg] as a function of temperature is approximated by
h
g
= 2.20269 ×10
7
+ 3.65317 × 10
5
T 2.25837×10
3
T
2
+ 7.3742 T
3
1.33437×10
2
T
4
+ 1.26913 × 10
5
T
5
4.9688 × 10
9
T
6
. (B.1)
We also use the temperature dependent steam viscosity
µ
g
= 5.46807 ×10
4
+ 6.89490 × 10
6
T 3.39999×10
8
T
2
+ 8.29842 × 10
11
T
3
9.97060×10
14
T
4
+ 4.71914 × 10
17
T
5
. (B.2)
The temperature dependent water viscosity
µ
w
is approximated by
µ
w
= 0.0123274+
27.1038
T
23527.5
T
2
+
1.01425×10
7
T
3
2.17342×10
9
T
4
+
1.86935×10
11
T
5
.
(B.3)
We assume that the steam is a ideal gas, so the steam density is a function of temper-
ature:
ρ
g
(T)=p
M
H
2
O
R
1
T
, (B.4)
where M
H
2
O
is the water molecular mass [Kg/m
3
], p is the pressure atmospheric [Pa]
and R=8.31 [J/mol K]. The quantity pM
H
2
O
/R is a constant.
158
The liquid water density is constant, and the value is 998.2Kg/m
3
.
We define
ˆ
H
r
and the water enthalpy per mass unit h
w
respectively as:
ˆ
H
r
(T)=(1
ϕ
)/
ϕ
C
r
T and h
w
= C
w
T/
ρ
w
. (B.5)
B.2 Constitutive relations
The relative permeability functions k
rg
and k
rw
are considered to be power functions
of their respective saturations i.e.
k
rg
=
s
g
1 s
wc
n
g
and k
rw
=
s
w
s
wc
1s
wc
n
w
for s
w
s
wc
,
0 for 0
s
w
s
wc
,
. (B.6)
For the computations we take n
w
= 4, = n
g
= 2. The connate water saturation s
wc
is
given in Table 2 below.
Table 2, Summary of physical input parameters and variables
Physical quantity Symbol Value Unit
Water, steam fractional functions f
w
, f
g
Eq. (5.6) . [m
3
/m
3
]
Porous rock permeability k 1.0 × 10
12
. [m
3
]
Water, steam relative permeabilities k
rw
,k
rg
Eq. (B.6) . [m
3
/m
3
]
Pressure p 1.0135×10
5
. [Pa]
Mass condensation rate q Eqs (5.1)-(5.2). [kg /(m
3
s)]
Water, steam phase velocity u
w
,u
g
Eq. (5.5) . [m
3
/(m
2
s)]
Total Darcy velocity u u
w
+ u
g
,Eq(5.7). [m
3
/(m
2
s)]
Water and rock heat capacity C
w
, C
r
4.22×10
6
, 2.029 × 10
6
. [J/(m
3
K)]
Steam and water enthalpies h
g
, h
w
Eqs. (B.1), (B.5.b). [J/m
3
]
Rock enthalpy H
r
(1
ϕ
)C
r
T. [J/m
3
]
Water, steam saturations s
w
,s
g
Dependent variables. [m
3
/m
3
]
Connate water saturation s
wc
0.15. [m
3
/m
3
]
Temperature T Dependent variable. [K]
Boiling point of water–steam T
b
373.15 . [K]
Water, steam thermal conductivity
κ
w
,
κ
g
0.652, 0.0208. [W/(mK)]
Rock, composite thermal conductivity
κ
r
,
κ
1.83 . [W/(mK)]
Water, steam viscosity
µ
w
,
µ
g
Eqs. (B.3) , (B.2). [Pa s]
Water, steam densities
ρ
w
,
ρ
g
998.2, Eq. (B.4) . [kg/m
3
]
Rock porosity (constant)
ϕ
0.38. [m
3
/m
3
]
B.3 Approximation for T + M
ρ
g
c
g
(T)/C
r
in the sr
As the temperature in the sr changes only from T
b
= 373,15K to around T = 400K,
from the expression for c
g
we can approximate T +(M
ρ
g
h
g
)/
ˆ
C
r
by a linear expression
159
aT + b (or, more generally, the by f(T)), where a e b are given in (B.8) .
f
(T)=aT + b. (B.7)
where
a
=
C
400
C
T
b
400 T
b
and b =
400C
T
b
T
b
C
400
400 T
b
, (B.8)
and
C
400
= 400 +
M
ρ
g
c
g
(400)
C
r
C
T
b
= T
b
+
M
ρ
g
c
g
(T
b
)
C
r
. (B.9)
370 375 380 385 390 395 40
0
0
0.5
1
1.5
2
2.5
x 10
−6
Normalizade error between the function f(T) and the function 1+(M
ρ
g
h
g
)/C
r
Temperature
Normalizade error
Solving (B.9) and (B.8) to find a and b we have
a
=
0.99952824228873 and b
=
0.28502504243858. (B.10)
Solving
(5.18) with linear denominator aT + b (see Appendix B) given by (B.7),we
have
u
= u
0
a
aT + b, (B.11)
where u
0
is the speed at the beginning of the rarefaction wave.
We use here
(B.7) in
λ
g
T
and
r
T
. After manipulations, we have
λ
g
T
=
u
ϕ
M
ρ
c
g
(T)
ˆ
C
r
(aT + b)
,
r
T
=
(1,
u
aT + b
) and c
g
(T)=
((
a 1)T + b)
ˆ
C
r
M
ρ
g
, (B.12)
160
where a e b are given in Eq. (B.7).
Finally, from
λ
T
and c
g
(T) given in (B.12) we have
λ
g
T
=
u
ϕ
(a 1)T + b
(aT + b)
. (B.13)
So from
(B.12) and (B.13), (B.16) can be written as
370 375 380 385 390 395 40
0
6
.2
6
.4
6
.6
6
.8
7
7
.2
7
.4
7
.6
7
.8
x 10
−4
λ
T
g
approximated graphic
Temperature
370 375 380 385 390 395 40
0
−2
0
2
4
6
8
10
12
x 10
−3
Relative
λ
T
g
approximated and real difference graphic
Temperature
Relative error
λ
g
T
·
r
T
=
u
ϕ
(a 1)T
(aT + b)
2
, (B.14)
Thus
λ
g
T
·
r
T
is negative for temperatures from 373.15K to 400K, so along rarefac-
tion curves defined by
(5.17.a) there are only decreasing temperatures; physically
this means that we are injecting steam at low temperature into the steam at higher
temperatures.
Thus the inflection locus in the plane
(T,u) consists of two parts
u
= 0 and T = 0, (B.15)
which are physically irrelevant.
B.3.1 Rarefaction wave behavior
Lemma B.3.1. If T decreases, so u decreases on rarefaction curve in direction where
the rarefaction exists.
Proof: As du
/dT satisfies (5.18. a),so
du
dT
> 0,
161
because all terms on the right hand side in (5.18. a) are positive.
So, if T increases then u increases; similarly, if T decreases then u decreases. From
Appendix B, equation
(5.4.3) T decreases along the rarefaction curve, so u decreases
on the rarefaction curve.
We study the rarefaction wave behavior. In this study, we find the region when
λ
T
increases and when it decreases along integral curves, and the inflection locus where
λ
T
is stationary. To do so, we need to calculate
λ
g
T
·
r
T
. (B.16)
We use
λ
g
T
and
r
T
given by (5.17.a) and (5.17.b), respectively, in Eq. (B.16):
λ
g
T
·
r
T
=
u
ϕ
M
ρ
g
ˆ
C
r
Tc
g
(T)
(
ˆ
C
r
T + M
ρ
g
c
g
(T))
2
, (B.17)
so in this case we also have that
λ
g
T
decreases when the temperature increases.
162
APPENDIX C
Applications
In Section C.1 we show that the formalism developed to class of equations (2.2) can
be applied to a hyperbolic system of form:
ˆ
G
(V)
t
+ u
ˆ
F
(V)
x
= 0, u = const. (C.1)
as a particular case. In Section C.2, we show an application to numerical method;
in Section C.3 we find the travelling waves independently on the variable u. We will
drop hats.
C.1 Hyperbolic waves in a physical situation
A very nice property of the formalism developed for shocks and rarefaction structures
of Eq.
(2.2) is that it can be applied to the hyperbolic system in the form (C.1),
illustrating that it is a particular system of the more general class
(2.2).
To do so, we consider the system
(C.1) with n state variables, where the cumulative
terms are G
=
(
G
1
, G
2
,···, G
n
)
and the flux terms are F =
(
F
1
, F
2
,···, F
n
)
.
Assume that there is no degeneracies. The RH condition for Eq.
(C.1) is:
v
s
[G
i
]=u[F
i
] for i = 1,2,··· ,n, (C.2)
where
[G
i
]=G
i
(V
+
) G
i
(V
) and [G
i
]=F
i
(V
+
) F
i
(V
). The RH locus satisfies:
[G
i
][ F
j
]=[G
j
][ F
i
] for all i, j = 1,2,··· ,n. (C.3)
If
[G
1
] = 0 and [F
1
] = 0, then the system of equations in (C.3) reduces to:
[G
1
][ F
j
]=[G
j
][ F
1
] for j = 2,3,···,n. (C.4)
163
The eigenvalues
λ
and right eigenvectors r of Eq. (C.1) are obtained by solving ( 2.30)
with A and B given by:
B
=
G
1
V
1
G
1
V
2
···
G
1
V
n
G
2
V
1
G
2
V
2
···
G
2
V
n
.
.
.
.
.
.
.
.
.
.
.
.
G
n
V
1
G
n
V
2
···
G
n
V
n
and A
=
u
F
1
V
1
u
F
1
V
2
··· u
F
1
V
n
u
F
2
V
1
u
F
2
V
2
··· u
F
2
V
n
.
.
.
.
.
.
.
.
.
.
.
.
u
F
n
V
1
u
F
n
V
2
··· u
F
n
V
n
.
We can rewrite
(C.1) in the form (2.2). To do so, we assume that u is the secondary
variable and V are the n primary variables. The system can be written as:
G
(V)
t
+
uF(V)
x
= 0, (C.5)
u
x
= 0, (C.6)
where Eq.
(C.1) represents the n first equations and (C.6) yields that u is constant in
the space. (Notice that in the classical hyperbolic equations u
1).
The matrix
M defined in (2.54) for the system (C.5)-(C.6) is:
[
G
1
]
F
+
1
F
1
[
G
2
]
F
+
2
F
2
.
.
.
.
.
.
.
.
.
[
G
n
]
F
+
n
F
n
0 11
. (C.7)
Assuming the same hypotheses of Corollary 2.3.1 and without loss of generality as-
suming that p
= 1 and q = 2, i.e., D
1
and D
2
are L.I., the RH locus are the solutions
of
det
[
G
1
]
F
+
1
F
1
[
G
2
]
F
+
2
F
2
[
G
i
]
F
+
i
F
i
= 0, (C.8)
for i
= 3,4,···,n + 1, where G
n+1
= 0 and F
n+1
= 1.
Notice that these are n
2 restrictions. For i = n + 1 in Eq. (C.8), we need to solve:
det
[
G
1
]
F
+
1
F
1
[
G
2
]
F
+
2
F
2
0 11
= 0,
that yields:
[G
2
][ F
1
]=[G
1
][ F
2
]. (C.9)
For i
= 3,4,···,n, we can write (C.8) as:
F
+
1
([G
2
]F
i
[G
i
]F
2
)+F
+
2
([G
i
]F
1
[G
1
]F
i
)+F
+
i
([G
1
]F
2
[G
2
]F
1
)=0 (C.10)
164
After some manipulations we obtain:
F
+
1
([G
2
][ F
i
] [G
i
][ F
2
]) + F
+
2
([G
i
][ F
1
] [G
1
][ F
i
]) + F
+
i
([G
1
][ F
2
] [G
2
][ F
1
]) = 0. (C.11)
Applying
(C.9) in Eq. (C.11), we obtain:
F
+
1
([G
2
][ F
i
] [G
i
][ F
2
]) + F
+
2
([G
i
][ F
1
] [G
1
][ F
i
]) = 0. (C.12)
(i) First we assume that [F
1
]=0. From Eq. ( C.12), it follows that
F
+
1
([G
2
][ F
i
] [G
i
][ F
2
]) F
+
2
[G
1
][ F
i
]=0. (C.13)
If
[F
2
] = 0 then [G
1
]=0. Since D
1
and D
2
are assumed to be L.I., then F
+
1
= 0 and for
all i
= 3,4,···,n:
[G
2
][ F
i
] [G
i
][ F
2
]=0. (C.14)
So
(C.9) together (C.14) is the RH locus for (C.5), (C.6) and coincides with the RH
locus of
(C.1).If[F
2
]=0, then it follows that for i = 3,4,··· ,n:
F
+
1
[G
2
] F
+
2
[G
1
]
[F
i
]=0. (C.15)
(ii ) Let us assume that [F
i
]=0 for all i = 3,4,··· ,n, then (C.4) is satisfied. If
F
+
1
[G
2
] F
+
2
[G
1
]=0, (C.9) we know that F
1
[G
2
] F
2
[G
1
]=0, implying that D
1
and
D
2
are linearly dependent and contradicting the hypotheses.
Finally assume that
[F
1
] = 0; substituting [G
2
] in (C.12) by [G
1
][ F
2
]/[F
1
] it follows
that:
F
+
1
[G
1
]
[
F
2
]
[F
1
]
[
F
i
] [G
i
][ F
2
]
+ F
+
2
(
[
G
i
][ F
1
] [G
1
][ F
i
]
)
= 0,
F
+
1
[F
2
]
[F
1
]
(
[
G
1
][ F
i
] [G
i
][ F
1
]
)
+ F
+
2
(
[
G
i
][ F
1
] [G
1
][ F
i
]
)
= 0,
(
[
G
1
][ F
i
] [G
i
][ F
1
]
)
[F
2
][ F
+
1
] F
+
2
[F
1
]
[F
1
]
= 0. (C.16)
If
[F
2
][ F
+
1
] F
+
2
[F
1
] = 0, then for all i = 3,4,··· ,n we obtain:
[F
i
][G
1
]=[G
i
][ F
1
],
that is the RH locus. On the other hand, if
[F
2
][ F
+
1
] F
+
2
[F
1
]=0, then:
F
+
2
F
1
F
2
F
+
1
= 0 (C.17)
Since
D
1
and D
2
are assumed to be LI, then:
det
[
G
1
]
F
+
1
[
G
2
]
F
+
2
= 0 or det
[
G
1
]
F
1
[
G
2
]
F
2
= 0 or det
F
1
F
+
1
F
2
F
+
2
= 0.
(C.18)
From
(C.17) and (C.9), we see that (C.18) cannot be satisfied. Thus we conclude that
the RH locus of
(C.1) coincides with the RH locus of the system (C.5), (C.6), defined
by Eq.
(C.8).
We have just proved the following Proposition:
165
Proposition C.1.1. Under the same hypotheses of Corollary 2.3.1, the RH locus of
(C.1) defined by Eq. ( C.4) coincides with the RH locus of the system ( C.5), (C.6).
Proposition C.1.2. The eigenvalues, the right and the left eigenvectors of the system
(C.5), (C.6) are
λ
=
ˆ
λ
, r =(
ˆ
r,0
) and =(,
n+1
), where
ˆ
λ
is the eigenvalue and
ˆ
r
and
ˆ
are the right and left eigenvectors of the system (C.1). The coordinate
n+1
is
determined by the system
(C.5), (C.6).
Proof: The eigenvalues of the system
(C.5), (C.6) are the solutions of:
det
u
F
1
V
1
λ
G
1
V
1
u
F
1
V
2
λ
G
1
V
2
··· u
F
1
V
n
λ
G
1
V
n
F
1
u
F
2
V
1
λ
G
2
V
1
u
F
2
V
2
λ
G
2
V
2
··· u
F
2
V
n
λ
G
2
V
n
F
2
.
.
.
.
.
.
.
.
.
.
.
.
u
F
n
V
1
λ
G
n
V
1
u
F
n
V
2
λ
G
n
V
2
··· u
F
n
V
n
λ
G
n
V
n
F
n
00··· 01
= 0, (C.19)
which reduces to:
det
u
F
1
V
1
λ
G
1
V
1
u
F
1
V
2
λ
G
1
V
2
··· u
F
1
V
n
λ
G
1
V
n
u
F
2
V
1
λ
G
2
V
1
u
F
2
V
2
λ
G
2
V
2
··· u
F
2
V
n
λ
G
2
V
n
.
.
.
.
.
.
.
.
.
.
.
.
u
F
n
V
1
λ
G
n
V
1
u
F
n
V
2
λ
G
n
V
2
··· u
F
n
V
n
λ
G
n
V
n
= 0, (C.20)
which yields the eigenvalues of the system
(C.1). The right and left eigenvectors
satisfy
(C.19), with r =(
ˆ
r,0
) and =(
ˆ
,
n+1
), where
ˆ
λ
is the eigenvalue and
ˆ
r and
ˆ
are the right and left eigenvectors of the system (C.1) and
n+1
=
ˆ
1
ˆ
F
1
+
ˆ
1
F
2
+ ···+
ˆ
n
F
n
,
where
ˆ
i
for i = 1,2,··· ,n are the coordinates of left eigenvector
ˆ
.
From Propositions C.1.1 and C.1.2 we obtain:
Corollary C.1.1. The wave structures for
(C.1) coincide with the wave structures for
(C.5), (C.6) .
C.2 Numerical Method
We propose a numerical method to solve Eq. (2.2) for all contiguous physical situa-
tions. Because the speed u does not appear in G
(V), there is an infinite speed mode
associated to u. Therefore the usual explicit finite difference numerical methods for
solving hyperbolic equations cannot be applied to Eq. (2.2
). Bruining et. al. have
applied In
[40], we have applied a similar method to the Riemann problem of nitrogen
and steam injection.
166
C.2.1 General Formulation
We represent the spatial position of any quantity by k and the time by symbol n.We
denote the time interval by t and the spatial interval by x.
For the numerical approximation, we use
V
k,n
= V(kx, nt) and u
k,n
= u(kx , nt), k Z, n N,
here kx and nt are, respectively, the position and the time where V and u are
evaluated.
We denote the variables to be found at time n
+ 1 and position k by
V
= V(kx, (n + 1)t) and u = u(kx,(n + 1)t). (C.21)
For simplicity, we abbreviate the known quantities
G
(V
k,n
)=G
k,n
, F(V
k,n
)=F
k,n
, G = G(V) and F = F(V). (C.22)
C.2.2 Numerical Method for the System
Using the notation of Section C.2.1, we propose an implicit method to approximate
Eq. ( 2.2) given by the following difference scheme:
G
G
k,n
t
=
uF + u
k1,n+1
F
k1,n+1
x
. (C.23)
where u and V are the unknowns defined in
(C.21). These equations are utilized to
find the unknown V and u at time n
+ 1 and position k.
Proposition C.2.1. If there exists a j such that F
j
(V) = 0 for each V in the domain,
the variables V can be obtained from the implicit scheme
(C.23) without calculating u.
Proof: We introduce the new variable
ˆ
u given by:
ˆ
u
=
ut
x
. (C.24)
So the system
(C.23) can be rewritten as:
G
(V) G
k,n
=
ˆ
uF
+
ˆ
u
k1,n+1
F
k1,n+1
. (C.25)
Since by hypothesis there exists a j such that F
j
(V) = 0, we can obtain u as function
of V as:
ˆ
u
=
G
k,n
j
G
j
(V)+
ˆ
u
k1,n+1
F
k1,n+1
j
F
j
. (C.26)
Notice that the equation C.26 is valid for some time n. Substituting
(C.26) in the
remaining equations of system
(C.25) we obtain for all i = j:
G
i
(V) G
k,n
i
F
j
=
G
k,n
j
G
j
(V)
F
i
+
ˆ
u
k1,n+1
F
k1,n+1
i
F
j
. (C.27)
167
Notice that we know u
0,n
for some time n, because we know the speed of injection
point. From Eq.
(C.26) we can obtain u
k1,n+1
as function of u
0,n+1
and V. After some
calculations we obtain:
ˆ
u
k1,n+1
=
k1
m
=1
G
m,n
j
G
m,n+ 1
j
(V)+
ˆ
u
0,n+1
F
0,n+1
j
F
k1,n+1
j
. (C.28)
Substituting
(C.28) in (C.27), we obtain for each i = j:
G
i
(V) G
k,n
i
F
j
F
k1,n+1
j
G
k,n
j
G
j
(V)
F
i
F
k1,n+1
j
k1
m=1
G
m,n
j
G
m,n+ 1
j
(V)
F
k1,n+1
i
F
j
ˆ
u
0,n+1
F
0,n+1
j
F
k1,n+1
i
F
j
= 0. (C.29)
Remark C.2.1. Since the matrix G/W is singular, it is not possible to use Newton
method to obtain
(V,u) in Eq. ( C.23). An advantage in utilizing the scheme (C.29) is
that u does not appear in this scheme, so we can linearize the reduced system or solve
it by a non-linear method, such as the Newton method.
C.3 Travelling waves and viscous profiles.
The system of equations (2.1) generically is obtained by simplifications from more
complex systems with diffusive terms of form:
t
G(V)+
x
u
F(V)=Q(V)+
x
B
V
x
, (C.30)
The m
×m + 1 matrix B := B(V) is called viscosity matrix; it can represent molecular
diffusion, capillarity effects, heat conduction and other physical phenomena.
The viscosity profile criterion is widely used for admissibility of physical shocks.
As a future work we will utilize this method to solve more complicated models where
elliptic regions appear, see
[24, 25, 26] and [61].
The travelling waves are solutions of
(C.30) that depend only on the parameter
ξ
= x v
S
t, where v
S
is the shock speed to be found. We rewrite this similarity
equation as:
W = W(x v
S
t), (C.31)
where we recall that
W =(V, u).
For the analysis of travelling waves and viscosity profile, we define the cumulative
condensation distribution
Q(x, t) and cumulative condensation Q
+
(t) by
Q(x, t)=
x
Q(x
,t)dx
, Q
+
(t)=
+
Q(x
,t)dx
. (C.32)
168
From Eq. (C.32), we can write Q = Q(x, t)/x. We also define Q
(t)=Q (x
,t) and
Q
+
(t)=Q (x
+
,t), where x
and x
+
are the points immediately on the left and right
of the transition between regions.
We can rewrite Eq.
(C.30) using the parameter
ξ
and we obtain:
v
s
dG
d
ξ
+
d(u F)
d
ξ
=
dQ
d
ξ
+
dB
d
ξ
dV
d
ξ
(C.33)
Integrating
(C.33) em
ξ
and setting that the limits satisfy:
lim
ξ
−→
(V(
ξ
),u(
ξ
),Q(V(
ξ
))) = (V
L
,u
L
,Q
L
), lim
ξ
−→
dV
d
ξ
= 0, (C.34)
lim
ξ
−→+
(V(
ξ
),u(
ξ
),Q(V(
ξ
))) = (V
R
,u
R
,Q
R
), lim
ξ
−→+
dV
d
ξ
= 0. (C.35)
We obtain the shock speed with
[Q]=Q
R
−Q
L
:
v
s
(G(V
R
) −G(V
L
)) + u
R
F(V
R
) u
L
F(V
L
)=[Q] , (C.36)
and the ordinary system of equations that determines the orbit of travelling wave
with
G
L
= G( V
L
) and F
L
= F(V
L
):
B
d(V(
ξ
))
d
ξ
= v
s
(G
L
−G(V(
ξ
))) + u(
ξ
)F(V(
ξ
)) u
L
F
L
(Q(V(
ξ
)) −Q
L
), (C.37)
d
Q
d
ξ
= Q. (C.38)
Lemma C.3.1. Applying E
∈Edefined in (2.17), (C.36) reduces to the RH condition
(2.50).
Lemma C.3.2. If there exists an i such that
F
i
(V) = 0 for each V (see Remark 2.2.3)
along the orbit of the system
(C.37), then in the variables V this orbit does not depend
explicitly on u.
Proof: The i-th line of the system
(C.37) is:
B
i
d(V(
ξ
))
d
ξ
= v
s
(G
L
i
−G
i
(V(
ξ
))) + u(
ξ
)F
i
(V(
ξ
)) u
L
F
L
i
(Q
i
(V(
ξ
)) −Q
L
i
), (C.39)
where
B
i
represents the i-th line of the matrix B. Assume that F
i
= 0 for some i,we
can obtain the Darcy speed as:
u
(
ξ
)=
u
L
F
L
i
B
i
d(V(
ξ
))/d
ξ
v
s
(G
L
i
−G
i
(V(
ξ
))) (Q
i
(V(
ξ
)) −Q
L
i
)
F
i
(V(
ξ
))
, (C.40)
169
After same calculations, for fixed i and k = 1,2,···,(i 1),(i + 1), ···,m, we can write
(C.37) as:
d
B
i
d
ξ
F
k
(V(
ξ
))
dB
k,
d
ξ
F
i
(V(
ξ
)) = u
L
F
L
i
F
k
(V(
ξ
)) −F
L
i
F
k
(V(
ξ
))
+
+
v
s

G
L
i
−G
i
(V(
ξ
))F
k
(V(
ξ
))
G
L
k
−G
k
(V(
ξ
))F
i
(V(
ξ
))

+
+(Q
k
(V(
ξ
)) −Q
L
k
)F
i
(V(
ξ
)) (Q
i
(V(
ξ
)) −Q
L
i
)F
k
(V(
ξ
)).
(C.41)
Since the number of equations and variables is the same, the proposition is proved.
Corollary C.3.1. Within each physical situation it is possible to apply E ∈E. In this
case, the orbits of the system
(C.41) can be calculated only in the space of primary
variables.
170
Livros Grátis
( http://www.livrosgratis.com.br )
Milhares de Livros para Download:
Baixar livros de Administração
Baixar livros de Agronomia
Baixar livros de Arquitetura
Baixar livros de Artes
Baixar livros de Astronomia
Baixar livros de Biologia Geral
Baixar livros de Ciência da Computação
Baixar livros de Ciência da Informação
Baixar livros de Ciência Política
Baixar livros de Ciências da Saúde
Baixar livros de Comunicação
Baixar livros do Conselho Nacional de Educação - CNE
Baixar livros de Defesa civil
Baixar livros de Direito
Baixar livros de Direitos humanos
Baixar livros de Economia
Baixar livros de Economia Doméstica
Baixar livros de Educação
Baixar livros de Educação - Trânsito
Baixar livros de Educação Física
Baixar livros de Engenharia Aeroespacial
Baixar livros de Farmácia
Baixar livros de Filosofia
Baixar livros de Física
Baixar livros de Geociências
Baixar livros de Geografia
Baixar livros de História
Baixar livros de Línguas
Baixar livros de Literatura
Baixar livros de Literatura de Cordel
Baixar livros de Literatura Infantil
Baixar livros de Matemática
Baixar livros de Medicina
Baixar livros de Medicina Veterinária
Baixar livros de Meio Ambiente
Baixar livros de Meteorologia
Baixar Monografias e TCC
Baixar livros Multidisciplinar
Baixar livros de Música
Baixar livros de Psicologia
Baixar livros de Química
Baixar livros de Saúde Coletiva
Baixar livros de Serviço Social
Baixar livros de Sociologia
Baixar livros de Teologia
Baixar livros de Trabalho
Baixar livros de Turismo