Transcript
Page 1: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

University of Groningen

Synthesis and evaluation of novel linear and branched polyacrylamides for enhanced oilrecoveryWever, Diego-Armando Zacarias

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

Document VersionPublisher's PDF, also known as Version of record

Publication date:2013

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):Wever, D-A. Z. (2013). Synthesis and evaluation of novel linear and branched polyacrylamides forenhanced oil recovery. Groningen: s.n.

CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

Download date: 11-10-2020

Page 2: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

179

Chapter 8

Towards new polymers for enhanced

oil recovery

Abstract

The progress booked in this project is discussed in terms of the problems

that have been overcome. Control of the acrylamide (AM) polymerization was

accomplished and this allows the preparation of polyacrylamide (PAM) with

variations in its chemical structure and molecular architecture. Branching in

PAM is presented as a new tool to significantly improve the solution viscosity

without changing the chemical structure of the polymer.

The non-ionic nature of the branched PAM renders it insensitive to the

presence of salt. More specifically, the solution viscosity and the elastic

response are not affected by the presence of salt. The rheological properties

of aqueous solutions are maintained when increasing the amount of salt

whereas those of the commercial HPAM are dependent on the salt

concentration.

The AM units in HPAM are also highly susceptible to alkaline hydrolysis at

high temperatures (T > 50 °C). N,N-dimethyl acrylamide (DMA), as a

hydrolysis resistant moiety, is a viable option especially in combination with

the increase thickening efficiency through branching. The oil recovery

efficiency of the branched PDMA polymers is similar to branched PAMs, as

evaluated using a 2D flow-cell that simulates residual oil trapped in dead-

ends. This makes them interesting polymers for application in EOR where

alkaline agents are used to in situ generate surfactants.

The obtained results constitute a breakthrough in the general field of

chemical product design for EOR. However, a further refinement of the used

criteria can be envisaged on the basis of the changing legislation (e.g.

currently in Norway). This implies the use of biologically degradable polymers

for underground injection as fitting a general “sustainability” concept for this

application. In this context, the potential of different biopolymers for EOR

where synthetic polymers cannot be used is briefly discussed. All the

currently investigated biopolymers are based on polysaccharides, albeit with

slightly different molecular structures. The thickening capability and

Page 3: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

180

resistance towards salt and temperature of the biopolymers are a strong

function of the molecular weight and their ability to form helical structures in

aqueous solutions. The elasticity of the biopolymer solutions are a strong

function of the molecular weight of the biopolymer, with the higher molecular

weight ones displaying a more pronounced elastic response.

Page 4: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

181

8.1. Introduction

Water soluble polymers in EOR have been investigated and applied in

different techniques, i.e. Polymer Flooding, Alkaline Surfactant Polymer (ASP)

Flooding, and Surfactant Polymer (SP) Flooding.1 When water soluble

polymers are applied in any of the former techniques, several different

aspects have to be considered. In the context of this dissertation, i.e. EOR,

the important ones (amongst others) are: the solution viscosity as a function

of the polymer concentration, the dependency of the solution viscosity on the

presence of salt (mono- and divalent ions), and the resistance (in terms of

solution viscosity) towards higher temperatures (T > 50 °C).

The dependency of the solution viscosity on the polymer concentration is

in general well known for homopolymers. The rheological properties depend

on the concentration regime in which the solution is. In general four different

concentration regimes can be distinguished (Figure 8.1).

Figure 8.1: Concentration regimes in polymeric solutions

At low polymer concentration, the polymer coils do not overlap and the

solution rheology can be described2 using Einstein’s equation3 for dilute

solutions of spherical particles:

(8.1)

where = viscosity of the solution, = viscosity of the solvent and = the

volume fraction of the particles.

At these low concentrations, the solution viscosity depends solely on the

volume fraction of the spheres, i.e. the polymer coils, and not on their size.

The polymer concentration at which the polymer coils start to touch (overlap)

each other is defined as the first critical overlap concentration (C*).

According to several estimations, the values for C* generally range between

Page 5: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

182

0.1 up to 5.0 wt.%.3 When the polymer concentration is increased beyond

C*, overlapping of the polymer coils becomes more prominent and the

solution viscosity increases significantly. The concentration regime starting

from the C* up to a second critical concentration (C**, above which a gel is

formed) is defined as the semi-dilute regime. The viscosity of a polymeric

solution in this regime is governed by the relaxation, i.e. reptation4, of the

entanglements in response to disturbances caused by deformation forces

(stresses). At higher concentrations (c > C**, polymer concentration [p] ≥

50 wt.%3), the rheological properties resemble those observed for polymers

in the melt state.3, 5, 6 However, when making allowances for the desired

application (i.e. EOR), the increase in the polymer concentration can lead to

significant problems in the reservoir. The propagation of polymer coils

through narrow pore throats “presses” the coils closer to each other. If the

coils are large enough, bridging (Figure 8.2) can arise which leads to

blockage of pores.7-10

Figure 8.2: Bridging in porous media10

An increase in the polymer concentration is also detrimental for the

economics of a flooding project given the scale of such projects (e.g. for the

Marmul field pilot-project ~25 ton/day of dry polymer has been used11, 12).

The molecular weight also affects the rheological properties of water

soluble polymers. In general, the solution viscosity increases with the

molecular weight of the polymer, and the dependency can be described using

the reptation model of de Gennes.4 At equal polymer concentration, an

Page 6: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

183

increase in the molecular weight will lead to an increase in the overlap

density, which in turn leads to longer relaxation times (synonymous of higher

viscosities). Nevertheless, the molecular weight of the polymers cannot be

increased indefinitely without leading to problems. Sensitivity towards

mechanical degradation becomes a significant problem as the molecular

weight of the polymers increases.13-17 In addition, the aforementioned

problem with bridging will be augmented with an increase in the molecular

weight.8, 9 In practice, high molecular weight HPAM leads to filter cake

formation on the surface (face plugging) of cores, especially low permeable

ones, and cannot be applied in a reservoir field with similar rock properties.18

The introduction of charges (in PAM, the hydrolysis degree) in a polymer

backbone will lead to an increase in the solution viscosity compared to its

uncharged analogue.19 The higher the amount of charged moieties, the more

stretched the polymer coil will be and thus the higher the solution viscosity.

This enables the use of lower molecular weight polymers without jeopardizing

the thickening capability. However, for the polymer to remain insensitive

towards salinity and hardness of the brine, the hydrolysis degree cannot be

too high (e.g. higher than 40 mol%).20 Therefore, although beneficial for the

thickening capability in de-ionized water, the presence of charged moieties

will result in sensitivity issues towards electrolytes. In the presence of

divalent ions (such as Ca2+), even precipitation can arise, due to inter-chain

complexation21, eventually leading to a complete loss of solution viscosity. In

practice the hydrolysis degree is fixed at 30 mol%. Nevertheless, to date

partially hydrolyzed polyacrylamide (HPAM) is the polymer of choice for

chemical EOR, mainly in connection to its relatively low price (2 - 4 €/kg).

This project started with identifying the limitations of the currently used

HPAM in enhanced oil recovery (vida supra). The main objective of this

dissertation was to tackle a couple of these limitations and present new

solutions. Firstly the controlled polymerization of acrylamide (AM) was

accomplished through the use of atomic transfer radical polymerization

(ATRP) in water. In addition, the “living” character of the polymerization

process offered the possibility of adding a second block of N-

isopropylacrylamide (NIPAM). The controlled polymerization of AM enabled

the design of different molecular architectures of polyacrylamide (PAM).

Subsequently evidence for the increased thickening capability of branched

PAM versus linear PAM could be provided. In addition, the presence of

branches (N > 8) increased the elastic response of aqueous solutions of the

polymers. Given the uncharged nature of the branched PAMs (compared to

the commercial HPAM with ~ 30 mol% of charged moieties), the presence of

Page 7: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

184

salt does not influence the solution properties making these polymers

particularly suitable for high salinity reservoirs.

Another limitation related to HPAM is the temperature stability. The use

of copolymers of AM and NIPAM provides polymers that display thermo-

thickening up to 80 °C, and are therefore resistant (in terms of solution

viscosity) to higher temperatures (T > 50 °C). In addition, the oil recovery

efficiency at high temperatures using the thermo-responsive copolymers is

significantly improved compared to a branched PAM analogue.

In the following sections, the branched PAMs prepared in this thesis are

discussed in terms of their rheological properties compared to either linear

PAMs or commercial linear HPAMs. Unresolved problems related to the use of

AM as a monomer are discussed and preliminary results on improvements are

presented. In addition, new preliminary results for other acrylamide-based

materials as well as several different biopolymers are presented in terms of

rheological behavior and oil recovery performance.

8.2. Thickening capability, comb-shaped PAM

The thickening capability of the currently used HPAM is due to its high to

ultra-high molecular weight (3.5 – 20·106 g/mol) and the presence of

charged (25-35 mol%) moieties.19, 22 According to the general theory of

polyelectrolyte solutions3, the presence of the charged moieties leads to

electrostatic repulsions and subsequently to prominent chain stretching.19

However, when dissolved in salt solutions the thickening capability is

significantly hampered (due to the electrostatic screening of the charged

moieties). Other ways of increasing the thickening capability of a polymer is

the introduction of hydrophobic moieties that will lead to aggregate

formation.19, 23 In this thesis (chapter 3 & 4, Figure 8.3), a new approach to

improve the thickening capability (in water solutions) of a polymer has been

developed.

The thickening capabilities of the branched PAMs depend on the

functionalization (number of arms) degree. A low number of arms (N ≤ 8)

leads to polymers which display a lower solution viscosity compared to linear

PAMs of equal theoretical overall molecular weight. This is attributed to the

inherent lower hydrodynamic volume of branched polymers.24, 25 For larger

number of arms (N ≥ 12), a higher solution viscosity is found when compared

to linear analogues. The branched PAMs with a relatively high number of

arms (N = 12, 13 and 17) possessed a higher hydrodynamic radius compared

to the branched PAMs with a low (N = 4 and 8) number of arms at equal total

molecular weight.

Page 8: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

185

Figure 8.3: New approach to increase the solution viscosity of aqueous solutions

In Chapter 1, the thickening capabilities of several different water soluble

polymers were plotted against each other in Figure 1.11. With the results of

Chapter 3 & 4 a comparison of the branched PAM with other (non-

hydrophobic) AM based polymers is performed and the results are displayed

in Figure 8.4.

As can be observed in Figure 8.4 A, the thickening capability of the HPAM

is the highest of the three included in the comparison. Remarkably, the

thickening capability of the branched PAM (with a lower molecular weight

than that of the linear PAM) is seven times as high as that of the linear PAM.

This demonstrates that the molecular architecture is a strong tool to improve

the thickening capabilities of water-soluble polymers in the concentration

regime useful for EOR. The thickening capabilities of the branched PAMs have

been extensively discussed in Chapters 3 & 4. Here the focus will be on the

salt resistance of the branched PAMs in terms of solution viscosity and the

viscoelastic response of aqueous solutions containing them.

Page 9: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

186

4,5

3,2

3,5

PAM HPAM 13-arm PAM

0,0

0,2

0,4

0,6

0,8

1,0

1,2 Viscosity

Vis

co

sity (

Pa

.s)

0

1

2

3

4

5

Molecular weightA

Mo

lecu

lar

we

igh

t (x

10

6 g

/mo

l)

3,2 3,2 3,2 3,2 3,2

3,5

HPAM

0,5 N

aCl

2,0 N

aCl

3,0 N

aCl

12,0 N

aCl

13-arm

PAM

0,0

0,1

0,2

0,3

0,4

0,5 Viscosity

Vis

cosity (

Pa.s

)

B

0

1

2

3

4

5

Molecular weight

Mo

lecula

r w

eig

ht (x

10

6 g

/mol)

Figure 8.4: Thickening abilities of different AM-based polymers, (A) the solution

viscosity (at = 10 s-1) of the polymer solution (1 wt.%) with corresponding molecular

weight and (B) the solution viscosity (at = 10 s-1) of the polymer solution (0.5 wt.%)

with corresponding molecular weight at different salt (NaCl) concentration for HPAM

and a 13-arm branched PAM (no salt)

8.3. Salt resistance, comb-shaped PAM

The salt sensitivity of HPAM is a well know problem given its ionic

character.1 The solution viscosity decreases significantly as the salt

concentration increases. Given that in all oil reservoirs brine (salt water) is

used, it is not a problem that can be circumvented by using deionized water.

In addition, in the presence of weak bases (such as sodium carbonate)

hydrolysis of the acrylamide moieties occurs which becomes extensive at

elevated temperatures (T > 60 °C) The injection of non-hydrolyzed PAM,

Page 10: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

187

rather than HPAM, has been proposed as a new method for EOR.26 The non-

hydrolyzed PAM will be hydrolyzed in-situ and the viscosity of the solution will

increase. For oil reservoirs where a high amount of salt is present the use of

pristine PAM can represent a good option. However, the low thickening

capability of linear PAM compared to linear HPAM will be detrimental for a

project given the higher amount of linear PAM required to match the viscosity

of the aqueous phase to that of the oil. Therefore, we propose the use of

branched PAM with its better thickening capability compared to linear PAM for

high salinity applications (Figure 8.5).

0 1 2 3 11 1210

1

102

103

A PK30-g

13-(PAM49225)

PK30-g13

-(PAM23810)

Poly(AM88630-ran-AA37470)

Poly(AM56135-ran-AA23730)

Poly(AM31515-ran-AA13320)

Vis

cosity (

mP

a.s

)

Concentration NaCl (wt. %)

0,1 1 10 100 1000

101

102

Vis

cosity (

mP

a.s

)

Shear rate (s-1)

PK30-g13

-(PAM49225)

Poly(AM88630-ran-AA37470)

PK30-g13

-(PAM23810)

Poly(AM31515-ran-AA13320)

B

Figure 8.5: A; the solution viscosity ( = 10 s-1) as a function of the salt (NaCl)

concentration for HPAM and branched PAM, and B; the viscosity functions of 2 charged

HPAMs and 2 uncharged branched PAMs

Page 11: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

188

As evident in Figure 8.5 A, the solution viscosity of the uncharged branched

PAMs (PK30-g13-(PAM23810) and PK30-g13-(PAM49225)) is not affected by

the presence of salt (up to 12 wt.% of NaCl). The solution viscosities ( = 10

s-1, [p] = 5000 ppm) of the charged linear HPAM are all higher than the

branched PAMs in de-ionized water. However, as the amount of salt increases

the solution viscosities of the branched PAMs remain constant while that of

the charged HPAMs decreases significantly. Remarkably, the solution

viscosity of a charged HPAM with a molecular weight between 8 – 10 · 106

g/mol decreases to values lower than that of the PK30-g13-(PAM49225) (Mn ≈

3.5 · 106 g/mol). This demonstrates the suitability in terms of the solution

viscosity of the branched PAMs for application in high salinity environments.

The shear thinning behavior of the aqueous solutions has also been probed.

As can be observed in Figure 8.5 B, this pseudoplasticity of the branched PAM

is similar to that of the charged HPAM with a molecular weight either 2 or 3

times as high as that of the branched PAM. In actual applications the

pseudoplastic behavior is preferred, given that a low viscosity at high shear

rates will require less pumping energy.

Another important parameter identified for an efficient oil recovery is the

viscoelasticity of the aqueous phase.27-34 In Figure 8.6, the viscoelastic

response of aqueous solutions containing either a linear HPAM or a branched

PAM is displayed.

As can be observed in Figure 8.6 A, the viscoelastic response of the

HPAM solution is dependent on the salt concentration. A significant decrease

in the elasticity of the solution can be clearly distinguished as the

concentration of the salt increases. The reduction19 in the hydrodynamic

volume of the polymer coils, due to electrostatic screening, is the accepted

explanation of the observed behaviour.3 The effective size of the polymer in

solution is smaller, and therefore the extent of overlapping is suppressed

which leads to a lower elastic response.

The results for the uncharged branched PAM (Figure 8.6 B) demonstrate

that the elastic response of the aqueous solution is not affected by the

presence of salt.

Page 12: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

189

0,1 1 10 100

0

30

40

50

60

70

80

90

Phase a

ngle

Frequency (rad/s)

NaCl concentration = 30000 ppm

NaCl concentration = 20000 ppm

NaCl concentration = 5000 ppm

NaCl concentration = 0 ppm

Increasing NaCl

concentration

A

0,1 1 10 100

0

30

40

50

60

70

80

90

B

Phase a

ngle

Frequency (rad/s)

NaCl concentration = 30000 ppm

NaCl concentration = 20000 ppm

NaCl concentration = 5000 ppm

NaCl concentration = 0 ppm

Figure 8.6: (A) the viscoelasticity as a function of the salt (NaCl) concentration for

HPAM (Mw = 3.2·106 g/mol, [p] = 1.0 wt.%), and (B) the viscoelasticity as a function

of the salt (NaCl) concentration of a branched PAM (Mw = 1.7·106 g/mol, [p] = 1.0

wt.%)

8.4. Hydrolysis resistance, comb-shaped PAM

The hydrolysis reaction of PAM is a well-known reaction that can be

catalysed either by an acid or a base.35 The hydrolysis reaction (Scheme 8.1)

leads to the formation of ammonia.

In ASP floods most often sodium carbonate is used as the alkali agent.

Therefore, the resistance to base catalysed hydrolysis of PAM is important. In

general there are two stages of the hydrolysis reaction.35 The first one (high

Page 13: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

190

rate) reaches hydrolysis degrees up to 40 mol% and is accelerated by

neighbouring carboxylate groups.

Scheme 8.1: Base catalysed hydrolysis of PAM

The second stage displays a ten times lower rate. This is suppressed by the

electrostatic repulsion between the carboxylate groups and the base, and the

increased viscosity due to chain stretching driven by electrostatic repulsion of

the carboxylate groups leads to mass transfer limitations. The parameters

that have been identified to accelerate the hydrolysis rate are high

temperatures, the presence of salts, polymer concentration, and high

base/polymer ratio.35-37 The characteristics of chemical EOR usually are a low

polymer concentration for economic reasons, temperatures above 50 °C

found for many oil reservoirs, and the presence of salts in the water used as

the displacing fluid. Therefore, it is obvious that the challenge to design a

polymer that can resist the base hydrolysis under the conditions in EOR is

important at an industrial level.

The use of other monomeric units that can withstand alkaline hydrolysis

is a viable option. Investigations towards novel multiblock co- and

terpolymers have demonstrated the effectiveness of changing the AM units

into other more resistant moieties.38, 39 Several different acrylamide based

monomers have been investigated as hydrolysis resistant ones (Figure

8.7).40-42

However, the homopolymers of DMA and AM display a markedly different

behaviour under the same conditions. After 50 hours, the hydrolysis degree

of poly(N,N-dimethylacrylamide) (PDMA) is only 2 mol%, while that of PAM

reached a hydrolysis degree of 30 mol% after only 2 minutes.43, 44 The

reactivity of PAM towards alkaline hydrolysis is 500 times higher compared to

that of PDMA and PAAEE.43, 44 The synthesis of the polymers have all been

through free radical polymerization. In order to benefit from the improved

thickening capability of branched polymers compared with linear ones, the

controlled polymerization of the hydrolysis resistant monomers is required.

The controlled polymerization of DMA, NIPAM and AAE has been

demonstrated already.45-50

Page 14: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

191

Figure 8.7: Hydrolysis resistant acrylamide based monomers

8.4.1. Results and discussion

Macroinitiators. The synthesis of the macroinitiators was performed

according to the Paal-Knorr reaction (Scheme 8.2) of a halogenated primary

amine with aliphatic perfectly alternating polyketones. The carbonyl

conversion was determined using elemental analysis. The characterization of

the macroinitiators has been extensively investigated in Chapter 3 & 4 and

therefore will not be discussed here. The properties of the macroinitiators

used in the synthesis of branched PDMA are listed in Table 8.1.

Table 8.1: Properties of the macro-initiators

Polyketone sample (PK30-Cla) Elemental composition

(C : H : N, wt%) XCO (%)b Pyrrole

unitsc Mn,GPC PDI

PK30 (virgin) 67.0 : 8.4 : 0.0 - 0 2 797 1.74

PK30-Cl4 58.6 : 7.1 : 1.6 18.87 4 2 447 2.02

PK30-Cl8 64.0 : 7.9 : 3.3 37.21 8 2 244 2.01

PK30-Cl13 62.9 : 7.6 : 4.9 61.14 13 2 072 1.97

a. Number indicates the ethylene content (%)

b. The conversion of the carbonyl groups of the polyketone

c. Average number of pyrrole units per chain

Page 15: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

192

The obtained, chemically modified polyketones are used as macroinitiators in

the ATRP of DMA for the preparation of comb-shaped polymers with a

different number of side chains. The synthesis of linear and comb-like PDMA

was performed according to Scheme 8.2.

Scheme 8.2: Synthesis of (A) linear PDMA and (B) comb PDMA

Table 8.2: Characteristics of the linear and branched PDMAs

Architecture Entry [M]0:[I]0:[CuCl]0:

[Me6TREN]0

M/s1/s2a (wt:vol:vol);

T; Time (min) Conv (%) Mn,tot Mn,SPAN

Linear

1 22 919:1:1.5:1.5 1:5 ; 25 °C; 180 58.4 1 326 825 1 326 825

2 51 623:1:1.5:1.5 1:5 ; 25 °C; 180 58.2 2 978 320 2 978 320

3 88 515:1:3.0:3.0 1:5 ; 25 °C; 180 47.0 4 124 011 4 124 011

4-arm 4 79 351:1:3.0:3.0 1:4:1/10; 25 °C; 60 70.1 5 514 111 2 759 853

8-arm 5 79 213:1:3.0:3.0 1:4:1/10; 25 °C; 60 64.5 5 064 788 1 268 994

13-arm

6 19 969:1:1.5:1.5 1:6:1/6 ; 25 °C; 130 39.9 789 831b 124 310

7 49 905:1:1.5:1.5 1:5:1/10; 25 °C; 180 46.3 2 290 499 355 181

8 99 226:1:3.0:3.0 1:5:1/20; 25 °C; 150 49.9 4 908 300 757 920

9 200 000:1:3.0:3.0 1:5:1/40; 25 °C; 180 30.3 6 007 278 926 993

17-arm 10 100 030:1:1.5:3.0 1:4:1/20; 25 °C; 180 41.0 4 065 549 481 097

a. M/s1/s2 = Monomer / solvent 1 / solvent 2 = N,N-dimethylacrylamide / water / acetone

b. Mn,GPC = 771 300 g/mol and the PDI = 1.80 as determined by aqueous GPC

The ratio between the initiator (or the macroinitiator) and the monomer was

varied in order to synthesize linear and comb-shaped PDMA with different

Page 16: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

193

molecular weights. The linear polymers were prepared using MClPr as the

initiator while the comb PDMAs the polyketone based macroinitiators were

used. Table 8.2 lists the results for the different polymers prepared.

The polymerization of DMA in water at room temperature using the

polymerization process described in Chapter 2 & 3 allows for the preparation

of linear and branched PDMA with relatively low dispersity indices. The

rheological properties depend on the number of arms (Figure 8.8).

0,1 1 10 100 1000

102

103

A

Vis

co

sity (

mP

a.s

)

Shear rate (s-1)

PK30-g13

-(PDMA23105), entry 7

PDMA30045, entry 2

0,1 1 10 100 1000

102

103

B

Vis

co

sity (

mP

a.s

)

Shear rate (s-1)

PK30-g17

-(PDMA41010), entry 10

PDMA41600, entry 3

0,1 1 10 100 100010

1

102

Vis

co

sity (

mP

a.s

)

Shear rate (s-1)

PK30-g13

-(PDMA49515), entry 8

PK30-g8-(PDMA51090), entry 5

PK30-g4-(PDMA55625), entry 4

C

Figure 8.8: Viscosity functions of (A) linear and 13-arm branched PDMA of similar

Mn,tot,[p] = 2.0 wt%, (B) linear and 17-arm branched PDMA of similar Mn,tot,[p] = 2.0

wt% and (C) 4-arm, 8-arm and 13-arm branched PDMA, [p] = 1.0 wt%

Increasing the number of arms (from N = 4 to 17) leads to a higher solution

viscosity at equal polymer concentration and molecular weight, similar to the

results obtained for the branched PAMs (Chapter 4). This is evident from the

Page 17: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

194

comparison between a 13-arm (entry 7) and a 17-arm (entry 10) branched

PDMAs with their corresponding linear analogues (entries 2 and 3

respectively). The comparison between a 4, 8 and 13-arm branched PDMA

further demonstrates the effect of the number of branches on the solution

viscosity.

The hydrolysis resistance of the linear and branched PDMA were

investigated under conditions resembling those found in actual chemical EOR

(Figure 8.9).

0 25 50 75 100 125 150 175 200

100

150

200

250

300

350

400

Vis

co

sity r

ete

ntio

n (

%)

Time (hours)

PAM21445, NaCl

PAM21445, NaCl-CaCl2

PK30-g13

-(PAM23810), NaCl

PK30-g13

-(PAM23810), NaCl-CaCl2

PK30-g13

-(PDMA49515), NaCl

PK30-g13

-(PDMA49515), NaCl-CaCl2

A

0 25 50 75 100 125 150 175 200

0

2

4

6

8

10

12

14

95100

B

Hydro

lysis

deg

ree (

%)

Time (hours)

PAM21445, NaCl

PK30-g13

-(PAM23810), NaCl

Figure 8.9: (A) Solution viscosity (in percentages from the starting value) as a

function of hydrolysis time for a linear and a branched PAM and a branched PDMA,

[p]=5000 ppm, (B); the hydrolysis degree of the linear and branched PAM as a

function of temperature as measured by 13C-NMR-spectroscopy

Page 18: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

195

The aqueous solution used in EOR usually remains for extensive times in the

reservoir. Periods of several months onshore, and up to more than one year

offshore have been stated.22, 51 The most important parameter for

maintaining the success of the polymer flood is the solution viscosity of the

aqueous phase.

Upon hydrolysis charged groups are randomly introduced in the polymer.

This will lead to electrostatic repulsion3 thus increasing the hydrodynamic

volume of the coils in solution and this is synonymous to a higher solution

viscosity. The increase in the solution viscosity as a function of the hydrolysis

time is significantly more pronounced for the linear PAM (PAM21445) when

compared to the branched PAM (PK30-g13-(PAM23810) and PDMA (PK30-g13-

(PDMA48515). The solution viscosity increases by more than 300 % of the

original value, both with and without CaCl2. This is strong evidence that the

hydrolysis of the linear PAM is extensive (while that of the branched analogue

is not), since in the absence of salt the solution viscosity increases with an

increase in the hydrolysis degree (up to a limiting value).26 The differences in

the solution viscosities between the presence of NaCl or NaCl-CaCl2 suggest

that either the hydrolysis is suppressed by the presence of CaCl2 or part of

the polymer precipitates by complex formation with Ca2+.52 The relatively low

increase in the solution viscosity of the branched PAM suggests a lower

hydrolysis rate compared to the linear analogue. However, the increase in

solution viscosity with an increase in the hydrolysis rate not necessarily has

to be equal for both the linear and the branched PAM. Therefore, the direct

measurement of the hydrolysis degree (by 13C-NMR) was carried out for the

two samples (Figure 8.9 B). The increase in the hydrolysis degree is similar

during the first couple of hours. After 50 hours the hydrolysis degree of the

linear PAM surpasses that of the branched PAM indicating that the branched

PAM is more resistance to alkaline hydrolysis compared to the linear

analogue. As can be observed in Figure 8.9 A, the change in the solution

viscosity of the branched PDMA is limited. This is strong evidence that the

branched PDMA is resistant towards alkaline hydrolysis, which is in line with

earlier reports.44

Increasing the residence time under the harsh conditions and the salt

concentration leads to a significant increase in the solution viscosity (Figure

8.10) for the linear PAM.

Page 19: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

196

0 10 20 30 40 50 60

0

50

100

150

200

250

300

A

Vis

cosity r

ete

ntion

(%

)

Time (days)

PK30-g13

-(PAM35275), NaCl

PK30-g13

-(PAM35275), NaCl-CaCl2

PAM35705, NaCl

PAM35705, NaCl-CaCl2

PK30-g13

-(PDMA23105), NaCl

PK30-g13

-(PDMA23105), NaCl-CaCl2

PDMA30045, NaCl

PDMA30045, NaCl-CaCl2

0 10 20 30 40 50 60

0

40

50

60

70

80

90

100

110

Vis

cosity r

ete

ntion

(%

)

Time (days)

Poly(AM31515-co-AA13320), NaCl

Poly(AM31515-co-AA13320), NaCl-CaCl2

PK30-g13

-(PDMA23105), NaCl

PK30-g13

-(PDMA23105), NaCl-CaCl2

PDMA30045, NaCl

PDMA30045, NaCl-CaCl2 B

Figure 8.10: (A) Solution viscosity (in percentages from the starting value) as a

function of hydrolysis time for a linear ([p] = 5000 ppm) and a branched PAM ([p] =

4900 ppm) and a linear ([p] = 5900 ppm) and a branched PDMA ([p] = 4500 ppm), at

equal molar concentration, (B) Solution viscosity (in percentages from the starting

value) as a function of the hydrolysis time for a linear ([p] = 10000 ppm) and a

branched PDMA ([p] = 6500 ppm) and a commercial HPAM ([p] = 5500 ppm), at equal

starting solution viscosity (measured at = 10 s-1)

The branched PAM (with similar molecular weight) displays at first an

increase in the solution viscosity (albeit less pronounced compared to its

linear analogue) and decreases slowly to below the starting viscosity.

The hydrolysis degree of the 62 days samples was determined by 13C-

NMR as being 38 and 33 mol% for, respectively, the linear and the branched

PAM. The solution viscosity of the linear and branched PDMA is not

Page 20: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

197

significantly affected by the conditions applied, even after more than 60

days. This suggests that little, if any, hydrolysis takes place. This is

confirmed by 13C-NMR where no carboxylate units could be detected (i.e.

below the detection limit of 13C-NMR) for the 62 days samples. The presence of CaCl2 also affected the solution viscosity of the samples;

however the differences between the samples with CaCl2 and the ones

without were not large. Although the presence of CaCl2 did not significantly

affect the solution viscosity of the samples, precipitation was observed in the

case of PAM-based polymers (Figure 8.11).

Figure 8.11: Precipitation of the commercial HPAM (with CaCl2) sample after 42 days

The solutions of the linear and branched PDMA stayed clear even after 62

days in the oven, whereas the linear HPAM became more turbid. This

indicates the formation of large aggregates.

8.5. Oil recovery, 2D flow-cell

The efficiency of the branched hydrolysis resistant PDMA in recovering oil

out of dead-ends was evaluated using the flow-cell (Chapter 7). In addition,

the oil recovery at higher temperatures (i.e. T = 70 °C) using the thermo-

responsive block copolymers (Chapter 6) was also evaluated.

8.5.1. Oil recovery efficiency

The efficiency in recovering residual oil by branched PDMA (at room

temperature) and branched random copolymer of AM and NIPAM (at room

temperature and 70 °C) has been evaluated (Figure 8.12).

Page 21: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

198

[1] Brine (30000 ppm NaCl) [2] Poly(AM31515-ran-AA13320)

[p] = 8700 ppm

Residual oil recovery = 7.6 % (± 1.8)

[3] PK30-g13-(PDMA23105)

[p] = 8000 ppm

Residual oil recovery = 8.9 % (± 1.8)

[4] Water, RT

[5] PK30-g13-(PAM3275), RT

[p] = 11000 ppm

Residual oil recovery = 4.8 % (±1.9)

[6] PK30-g13-(PAM1405-ran-PNIPAM1405)

[p] = 9000 ppm, RT

Residual oil recovery = 3.6 % (±1.9)

[7] Water, 70 °C

[8] PK30-g13-(PAM3275), 70 °C

[p] = 11000 ppm

Residual oil recovery = 8.6 % (±1.8)

[9] PK30-g13-(PAM1405-ran-PNIPAM1405)

[p] = 9000 ppm, 70 °C

Residual oil recovery = 50.2 % (±1.0)

Figure 8.12: Oil recovery out of dead ends using branched PDMA ([3]) compared to

brine ([1]) and the commercial polymer ([2])at room temperature using crude oil, and

branched copolymers of AM and NIPAM ([6] and [9]) compared to water ([4] and

[7]) and branched PAM of similar molecular weight ([5] and [8]) both at room

temperature and at 70 °C using a mixture of crude oil and cyclo octane (2-1 vol.%)

The efficiency of the recovery of residual oil using the branched PDMA ([3]

PK30-g13-(PDMA23105)) is similar to that of the branched PAM (Chapter 7,

[D] PK30-g13-(PAM35275) and that of the commercial polymer ([2],

poly(AM31515-co-AA13320). However, the polymer concentration required to

match the solution viscosity of the water phase with that of the oil is higher

compared to a branched PAM of the same molecular weight. Nevertheless,

the ability to recover part of the residual oil makes these hydrolysis resistant

branched PMDA polymers potential candidates for EOR where alkali is also

used to generate in situ surfactants.

The residual oil recovery efficiency of the branched thermo-responsive

copolymers is slightly higher compared to a branched PAM ([5] PK30-g13-

(PAM3275)), similar molecular weight). However, when performing the

Page 22: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

199

comparison at 70 °C, different recovery efficiencies are observed for the

branched PAM and thermo-responsive copolymer. The recovery efficiency of

the branched PAM increases from 4.8 to 8.6 %. This can be attributed to the

improved mobility ratio (equation 1.3, Chapter 1) due to the lower viscosity

of the oil (the decrease in the oil viscosity is more pronounced that the

decrease of the polymer solution viscosity).

When comparing the recovery efficiency of the branched thermo-

responsive copolymer a significantly higher efficiency is observed. This

cannot reside only in the decrease of the oil viscosity at higher temperatures.

The higher oil recovery efficiency of the branched copolymer ([9] PK30-g13-

(PAM1405-co-PNIPAM1405)) at 70 °C is therefore attributed to the increased

solution viscosity (Chapter 6). The mobility ratio is lower than unity (and thus

lower at 70 °C compared to at RT) given the higher solution viscosity, and

thus a better displacement of the oil takes place. However, from a practical

point of view, the polymer concentration can be decreased until the solution

viscosity at 70 °C matches that of the oil.

The increase in oil recovery efficiency at higher temperatures makes

these types of copolymers interesting candidates for application in EOR where

the reservoir temperatures are high (i.e. T ≥ 50 °C).

8.6. Biopolymers for EOR

In certain regions of the world regulations stipulate that if a polymer is

used in recovering oil, it has to be reusable or biodegradable. If a synthetic

polymer is used, the produced mixture of oil and water (containing the

synthetic polymer) has to be separated and the water phase must be re-

injected. However, in most field application the polymer that is produced

along with the oil has been either chemically or thermally degraded and

therefore cannot be re-injected.22 Therefore, the use of biopolymers is almost

inevitable and a lot of effort has been put in developing biopolymers for EOR.

Although there are many examples of biopolymers that can be used for EOR,

only xanthan gum has been applied in actual oil reservoirs22, 53, although

there are current (pilot) projects under way with other water soluble

biopolymers, such as schizophyllan.

8.6.1. Thickening capability and viscoelasticity

Most of the biopolymers that have been considered so far for EOR are

polysaccharides.22 The ability of these type of polymers to increase the

viscosity of an aqueous solution is based on their high molecular weight and

in some cases the rigidity of the polymeric chains.19 Although there are many

Page 23: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

200

different types and sources of biopolymers, not all of them are soluble in cold

water. In many cases boiling water is required before complete dissolution of

the polymeric chains is obtained. From an economical and practical (remote

locations of many oil reservoirs) point of view, the dissolution in cold water is

preferred. Although only xanthan gum has been applied so far in chemical

EOR, there are many other different biopolymers that might be suitable. The

thickening capability54-56 of several different biopolymers is displayed in

Figure 8.13.

2

0,09

0,7

1,5

1,08

0,66

2,6

5

0,96

Xanthan g

um

Meth

yl ce

llulose

CM ce

llulose

(LM

)

CM ce

llulose

(HM

)

Chitosa

n

Lambda-c

arrageenan

Guar gum

Sclero

glucan

Schizo

phyllan

0

1

2

3

4

5

6

7

Viscosity

Vis

co

sity (

Pa

.s)

0

1

2

3

4

5

6

7

Molecular weight

Mo

lecu

lar

we

igh

t (x

10

6 g

/mo

l)

Figure 8.13: Thickening capabilities (viscosity measured at = 10 s-1) of different

biopolymers at a polymer concentration of 1 wt.%

As can be observed in Figure 8.12, there are several other biopolymers that

can significantly increase the viscosity of a water solution. Scleroglucan, a -

1,3 linked D-glucose with single D-glucose side chains linked -1,6 every

third unit57, has long been seen as a good substitute for xanthan gum58,

especially in oil reservoirs where high temperature and high salt

concentration (given the non-ionic character of scleroglucan) are found.59, 60

Another biopolymer that has been identified as a suitable biopolymer for

EOR is schizophyllan (chemically the same as scleroglucan).61 This resides

mainly in its ability to increase the solution viscosity even at very low

polymer concentration (i.e. a solution viscosity of 10 mPa.s, = 10 s-1 and a

[p] = 200 ppm).61 In addition, the solution viscosity of an aqueous solution

containing schizophyllan only decreases by 10 % when heated up to 130

Page 24: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

201

°C.61 By comparison, an aqueous solution of xanthan gum decreases by 95

%.61

Carboxy-methyl cellulose (CM cellulose) has also been considered as a

good candidate for EOR.53 The addition of carboxy-methyl groups to cellulose

makes the biopolymer soluble in cold water.53 This makes it attractive for

EOR since no specialty dissolution equipment is required. However, given the

ionic character of carboxy-methylcellulose, the solution viscosity is sensitive

to the presence of salt.53 Depending on the molecular weight (low or high [LM

or HM]) of the parent cellulose polymer a different thickening capability is

observed.

Chitosan has also been shown to increase the solution viscosity of a

water solution significantly.55 However, chitosan is only soluble in acidic

media62, 63 (i.e. pH62 < 6.0) which will significantly hamper its application in

EOR. Nevertheless, the high thickening capability of chitosan still makes it an

interesting polymer as a rheology modifier.

Methyl-cellulose, on the other hand, is soluble in cold neutral water.

Although it’s capability to increase the solution viscosity is less than most of

the biopolymers, its low molecular weight might make it suitable for low

permeable reservoirs. A peculiar behaviour of methyl-cellulose is its gelation

(in water solution) upon heating due to hydrophobic associations.64, 65 The

gelation is reversible; upon cooling the aqueous solution will return to its

original state.64, 65 Nevertheless, the hydrophobic character of parts of the

biopolymer might lead to enhanced adsorption (higher resistant factors,

indicating a higher layer thickness) on the rock surface similar to that

observed for hydrophobically modified polymers.23, 66

Guar gum has also been investigated for application in EOR. It is used

already to control the rheological properties of drilling muds. The thickening

capability of guar gum is higher than xanthan gum, but in solution of high

salinity guar gum is highly sensitive towards high temperatures limiting its

application.67

Little effort has been aimed at investigating the viscoelasticity of aqueous

solutions containing biopolymers. Experiments and mathematical models

have demonstrated the importance of the viscoelasticity of the solution on

the recovery of residual oil27-30, 30-34, although so far no consensus has been

reached. The viscoelasticity of some commercial biopolymers has been

evaluated and the results are displayed in Figure 8.14.

The results suggest that the elastic response of the aqueous solutions

increases as the molecular weight increase. This indicates that the extent of

overlapping is higher for the higher molecular weight polymers. Although the

Page 25: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

202

molecular weight of xanthan gum is not as high as that of CM cellulose, its

elastic response is much more pronounced. A possible explanation for this is

the rigidity of the polymeric chains. Xantham gum is known to form helices in

water solutions.68-71

0,1 1 10 100

10-4

10-3

10-2

10-1

100

101

102

A

G' G

" (P

a)

Frequency (rad/s)

Filled symbols = G'

Empty symbols = G"

= Xanthan gum

= Guar gum

= CM cellulose

= Methyl cellulose

= -Carrageenan

0,1 1 10 100

0

10

20

30

40

50

60

70

80

90

B

Ph

ase

an

gle

Frequency (rad/s)

Methyl cellulose

-Carrageenan

Guar gum

CM cellulose

Xanthan gum

Figure 8.14: (A), the loss and elastic modulus as a function of the frequency of

different biopolymers ([p]=1.0 wt.%) and (B), the phase angle as a function of the

frequency of the same biopolymers

8.7. Conclusion

Currently used partially hydrolyzed polyacrylamide (HPAM) in EOR has

several limitations. The main objective of this dissertation was to design new

acrylamide based polymers that provide solutions to the limitation of the

Page 26: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

203

aforementioned polymer. The first hurdle that had to be passed was the

controlled polymerization of AM. This was accomplished through the use of

atomic transfer radical polymerization (ATRP) in water at room temperature.

Furthermore, given the “living” character of the polymerization process a

second block of N-isopropylacrylamide (NIPAM) can be added to the first AM

block. With the accomplishment of controlled polymerization of AM, PAM with

different molecular architecture could be envisaged. This was achieved

through the use of functionalized (with halogens) alternating aliphatic

polyketones. Subsequently, evidence for the increased thickening capability

of bottle-brush PAM compared to a linear analogue was provided. The

presence of branches (N > 8) increased the elastic response of aqueous

solutions of the polymers. Also, given the uncharged nature of the bottle-

brush PAMs (compared to the commercial HPAM [~30 mol% charged

moieties]), the presence of salt does not influence the solution properties

making these polymers particularly suitable for high salinity reservoirs.

Another limitation related to HPAM is the temperature stability. Rheological

characterization demonstrated that copolymers of AM and NIPAM display

thermo-thickening behavior up to 80 °C, and are therefore resistant (in terms

of solution viscosity) to higher temperatures (T > 50 °C). The increased oil

recovery efficiency at high temperature (T = 70 °C) demonstrates the

potential of the thermo-responsive polymers for EOR.

Since in many cases alkaline agents are used in combination with

polymers, hydrolysis of the AM units in HPAM is extensive, especially at high

temperatures (T > 50 °C). The use of hydrolysis resistant moieties such as

DMA is promising, more so in combination with the increased thickening

capability through branching. The oil recovery efficiency of the branched

PDMA polymers is similar to branched PAMs, and these are therefore good

candidates for application in EOR where alkaline agents are used to generate

in situ surfactants.

The potential of using biopolymers for EOR where synthetic polymers

cannot be applied is briefly discussed. Most of the investigated biopolymers

are polysaccharides, with the differences being the source and molecular

structure of the polymeric chains. Their thickening capability is a function of

the molecular weight and the resistance towards salt depends on their ability

to form helical structures in aqueous solutions.

8.8. Acknowledgements

This work is part of the Research Program of the Dutch Polymer Institute

DPI, Eindhoven, The Netherlands, project #716.

Page 27: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

204

8.9. References

1. Thomas, A.; Gaillard, N.; Favero, C. Oil Gas Sci Technol 2012, 6, 887. 2. Guyot, A.; Chu, F.; Schneider, M.; Graillat, C.; McKenna, T. F. Progress in Polymer

Science 2002, 8, 1573. 3. Stokes, R. J.; Evans, D. F. Fundamentals of interfacial engineering; Wiley-VCH: New

York, 1997; . 4. Degennes, P. G. J. Chem. Phys. 1971, 2, 572. 5. Ferry, J. D. Viscoelastic properties of polymers; John Wiley & Sons: New York, 1980;

, pp 641. 6. Baumgartel, M.; Willenbacher, N. Rheologica Acta 1996, 2, 168. 7. Chauveteau, G.; Denys, K.; Zaitoun, A. SPE 2002, SPE-75183. 8. Zitha, P. L. J.; Botermans, C. W. SPE 1998, SPE-36665. 9. Zitha, P. L. J.; van Os, K. G. S.; Denys, K. F. J. SPE 1998, SPE-39675. 10. Zitha, P. L. J.; Chauveteau, G.; Leger, L. J. Colloid Interface Sci. 2001, 2, 269. 11. Shell Global Solutions International BV 2012. 12. Al-Saadi, F. S.; Amri, B. A.; Nofli, S.; Van Wunnik, J.; Jaspers, H. F.; Harthi, S.;

Shuaili, K.; Cherukupalli, P. K.; Chakravarthi, R. SPE 2012, SPE 154665,. 13. Moreno, R. A.; Muller, A. J.; Saez, A. E. Polymer Bulletin 1996, 5, 663. 14. Maerker, J. M. Society of Petroleum Engineers Journal 1975, 4, 311. 15. Maerker, J. M. Society of Petroleum Engineers Journal 1976, 4, 172. 16. Seright, R. S.; Seheult, M.; Talashek, T. SPE 2009, 5, SPE-115142-PA. 17. Zaitoun, A.; Makakou, P.; Blin, N.; Al-Maamari, R. S.; Al-Hashmi, A. R.; Abdel-

Goad, M.; Al-Sharji, H. H. Spe Journal 2012, 2, 335. 18. Wever, D. A. Z. 2009. 19. Wever, D. A. Z.; Picchioni, F.; Broekhuis, A. A. Prog. Polym. Sci. 2011, 1558. 20. Shupe, R. D. J. Petrol. Technol. 1981, 8, 1513. 21. Peng, S.; Wu, C. Macromolecules 1999, 3, 585. 22. Lake, L. W. Enhanced Oil Recovery; Prentice-Hall Inc.: Englewood Cliffs, NJ, 1989;

Vol. 1, pp 550. 23. Taylor, K. C.; Nasr-El-Din, H. A. J. Petrol. Sci. Eng. 1998, 3-4, 265. 24. Anthony, A. J.; King, P. H.; Randall, C. W. J Appl Polym Sci 1975, 1, 37. 25. Kulicke, W. M.; Horl, H. H. Colloid Polym. Sci. 1980, 7, 817. 26. Levitt, D. B.; Pope, G. A.; Jouenne, S. Spe Reservoir Evaluation & Engineering

2011, 3, 281. 27. Wang, D.; Cheng, J.; Yang, Q.; Gong, W.; Li, Q.; Chen, F. SPE 2000, SPE-63227-

MS. 28. Wang, D.; Xia, H.; Liu, Z.; Anda, Q.; Yang, Q. SPE 2001, SPE-68723-MS. 29. Yin, H.; Wang, D.; Zhong, H. SPE 2006, SPE-101950-MS. 30. Xia, H.; Wang, D.; Wang, G.; Wu, J. Petrol. Sci. Technol. 2008, 4, 398. 31. Hou, J. R.; Liu, Z. C.; Zhang, S. F.; Yue, X.; Yang, J. Z. Journal of Petroleum

Science and Engineering 2005, 3-4, 219. 32. Zhang, L.; Yue, X.; Guo, F. Pet. Sci. 2008, 1, 56. 33. Xia, H.; Ju, Y.; Kong, F.; Wu, J. SPE 2004, SPE-88456-MS.

34. Zhang, L.; Yue, X. J. Cent. South Univ. T. 2008, 84. 35. Kurenkov, V. F.; Hartan, H. G.; Lobanov, F. I. Russian Journal of Applied Chemistry

2001, 4, 543. 36. Maurer, J. J.; Harvey, G. D.; Klemann, L. P. Abstracts of Papers of the American

Chemical Society 1986, 152. 37. Went, P. M.; Evans, R.; Napper, D. H. Journal of Polymer Science Part C-Polymer

Symposium 1975, 49, 159. 38. Gaillard, N.; Giovannetti, B.; Favero, C. SPE 2010, SPE-129756. 39. Vermolen, E. C. M.; van Haasterecht, M. J. T.; Masalmeh, S. K.; Faber, M. J.;

Boersma, D. M.; Gruenenfelder, M. SPE 2011, SPE-141497-MS,. 40. Gelfi, C.; Debesi, P.; Alloni, A.; Righetti, P. G. J. Chromatogr. 1992, 1-2, 333. 41. Miertus, S.; Righetti, P. G.; Chiari, M. Electrophoresis 1994, 8-9, 1104.

Page 28: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

205

42. SimoAlfonso, E.; Gelfi, C.; Lucisano, M.; Righetti, P. G. Journal of Chromatography a 1996, 1-2, 255.

43. Righetti, P. G.; Gelfi, C. Anal. Biochem. 1997, 2, 195. 44. Righetti, P. G.; Chiari, M.; Nesi, M.; Caglio, S. J. Chromatogr. 1993, 2, 165. 45. Narumi, A.; Chen, Y.; Sone, M.; Fuchise, K.; Sakai, R.; Satoh, T.; Duan, Q.;

Kawaguchi, S.; Kakuchi, T. Macromolecular Chemistry and Physics 2009, 5, 349. 46. Neugebauer, D.; Matyjaszewski, K. Macromolecules 2003, 8, 2598. 47. Kim, K.; Kim, J.; Jo, W. Polymer 2005, 9, 2836. 48. Mueller, A. H. E.; Millard, P.; Mougin, N. C.; Boeker, A. Abstracts of Papers of the

American Chemical Society 2008. 49. Xia, Y.; Burke, N. A. D.; Stover, H. D. H. Macromolecules 2006, 6, 2275. 50. Xia, Y.; Yin, X. C.; Burke, N. A. D.; Stover, H. D. H. Macromolecules 2005, 14,

5937. 51. Sorbie, K. S. Polymer-improved oil recovery; CRC Press: Boca Raton, FL, 1991; . 52. Moradi-Araghi, A.; Doe, P. H. SPE 1987, SPE-13033-PA. 53. Donaldson, E. C.; Chilingarian, G. V.; Yen, T. F. Enhanced Oil Recovery II,

processes and operations; Elsevier: Amsterdam, The Netherlands, 1989; Vol. 2, pp 604.

54. Enomoto, H.; Einaga, Y.; Teramoto, A. Macromolecules 1984, 8, 1573. 55. Desbrieres, J. Biomacromolecules 2002, 2, 342. 56. Tipvarakarnkoon, T.; Senge, B. Ann T Nord Rheol Soc 2008, 16,. 57. Grassi, M.; Lapasin, R.; Pricl, S. Carbohydr. Polym. 1996, 2, 169. 58. Dixon, B. Bio-Technology 1985, 7, 601. 59. Kalpakci, B.; Jeans, Y. T.; Magri, N. F.; Padolewski, J. P. SPE 1990, SPE-20237-

MS,. 60. Kulawardana, E. U.; Koh, H.; Kim, D. H.; Liyanage, P. J.; Upamali, K. A. N.; Huh,

C.; Weerasooriya, U.; Pope, G. A. SPE 2012, SPE-154294-MS,. 61. Leonhardt, B. SPE 2011. 62. Pillai, C. K. S.; Paul, W.; Sharma, C. P. Progress in Polymer Science 2009, 7, 641. 63. Badawy, M. E. I.; Rabea, E. I. International Journal of Carbohydrate Chemistry

2011, 460381 (29 pp.). 64. Desbrieres, J.; Hirrien, M.; Ross-Murphy, S. B. Polymer 2000, 7, 2451. 65. Kobayashi, K.; Huang, C. I.; Lodge, T. P. Macromolecules 1999, 21, 7070. 66. Volpert, E.; Selb, J.; Candau, F.; Green, N.; Argillier, J. F.; Audibert, A. Langmuir

1998, 7, 1870. 67. Davison, P.; Mentzer, E. SPE 1982, SPE-9300-PA. 68. Dentini, M.; Crescenzi, V.; Blasi, D. Int. J. Biol. Macromol. 1984, 2, 93. 69. Holzwarth, G. Biochemistry 1976, 19, 4333.

70. Morris, E. R.; Rees, D. A.; Young, G.; Walkinshaw, M. D.; Darke, A. J. Mol. Biol. 1977, 1, 1.

71. Norton, I. T.; Goodall, D. M.; Frangou, S. A.; Morris, E. R.; Rees, D. A. J. Mol. Biol. 1984, 3, 371.

72. Parker, W. O.; Lezzi, A. Polymer 1993, 23, 4913. 73. Jones, D. M.; Walters, K. Rheologica Acta 1989, 6, 482.

Page 29: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

206

Appendix 8.A

8.A.1. Experimental section

Chemicals. N,N-dimethylacrylamide (DMA, ≥99%), copper(I) bromide

(CuBr, 98%), copper(I) chloride (CuCl, 98%), methyl 2-chloropropionate

(MeClPr, 97%), sodium chloride (NaCl, ≥99.5%), glacial acetic acid, ethanol,

and diethyl ether were purchased from Sigma Aldrich. Calcium chloride

dihydrate (CaCl2 · 2 H2O, 99%) and sodium bicarbonate (NaHCO3, ≥99%)

were purchased from Merck. CuBr & CuCl were purified by stirring in glacial

acetic acid for at least 5 hours, filtering, and washing with glacial acetic acid,

ethanol and diethyl ether (in that order) and then dried at reduced

pressure.46 All the other chemicals were reagent grade and used without

further purification.

Linear polymerization. A 250-mL three-necked flask was charged with

demineralized water and DMA. Subsequently, the mixture was degassed by

three freeze-pump-thaw cycles. A nitrogen atmosphere was maintained

throughout the remainder of the reaction steps. CuCl and the ligand

(Me6TREN) were then added to the flask and the mixture was stirred for 10

minutes. The flask was then placed in an oil bath at 25 °C. The reaction was

started by the addition of the initiator (MeClPr) using a syringe. After the

pre-set reaction time, the mixture was exposed to air and milli-Q water was

added. The contents were then purified via dialysis using membrane tubing

Spectra/Por® Dialysis Membrane (molecular weight cut off [MWCO] = 12 000

– 14 000 g/mol). The product was then dried in an oven at 65 °C until

constant weight and then ground.

Macroinitiators. The PK30 functionalization was performed according

(Scheme 8.A.1) to the published method. The reactions were performed in a

sealed 250 ml round bottom glass reactor with a reflux condenser, a U-type

anchor impeller, and an oil bath for heating.

For the preparation of PK30-Cl12 (taken here as an example) 3-

chloropropylamine hydrochloride (9.89 g, 53.6 mmol) was dissolved in

methanol (90 ml) to which an equimolar amount of sodium hydroxide (2.15

g, 53.6 mmol) was added. After the polyketone (10 g, 76 mmol of dicarbonyl

units) was preheated to the liquid state at the employed reaction

temperature (100 °C), the amine was added drop wise (with a drop funnel)

into the reactor in the first 20 min. The stirring speed was set at a constant

value of 500 RPM. During the reaction, the mixture of the reactants changed

from the slight yellowish, low viscous state, into a highly viscous brown

homogeneous paste. The product was dissolved in chloroform and afterwards

washed with demineralized water. The two phases (organic & water) were

Page 30: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

207

separated in a separatory funnel. The polymer was isolated by evaporating

the chloroform at reduced pressure at room temperature. The product, a

brown viscous paste (low degree of functionalization) or a brown powder

(high degree of functionalization), was finally freeze dried and stored at -18

°C until further use. Some properties of the macro-initiators are given in

Table 1. The macro-initiators were characterized using elemental analysis and 1H-NMR spectroscopy (in chloroform).

Scheme 8.A.1: Synthesis of the macro-initiators

The conversion of carbonyl groups of the polyketone was determined using

the following formula:

(8.A.1)

, is the average number of carbons in n-m (see Scheme 8.2)

, is the average number of carbons in m (see Scheme 8.2)

molecular weight of nitrogen

molecular weight of carbon

The number of pyrrole units was determined using the conversion of the

carbonyl groups of the polyketone and formula 2:

(8.A.2)

= the average molecular weight of the parent (unmodified) polyketone

= the average molecular weight of the repeating unit of polyketone

Page 31: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

208

Comb polymerization. A 250-mL three-necked flask was charged with the

macro-initiator. Enough acetone (typically 5-10 ml) was added to dissolve the

macro-initiator. Demineralized water and DMA were then added to the

solution. Subsequently, the mixture was degassed by three freeze-pump-

thaw cycles. A nitrogen atmosphere was maintained throughout the

remainder of the reaction steps. CuBr was then added to the flask and the

mixture stirred for 10 minutes. The flask was then placed in an oil bath at 25

°C. The reaction was started by the addition of the ligand (Me6TREN) using a

syringe. After the pre-set reaction time, the mixture was exposed to air and

the mixture was diluted with demineralized water. The reaction mixture was

then purified via dialysis using membrane tubing Spectra/Por® Dialysis

Membrane (molecular weight cut off [MWCO] = 12 000 – 14 000 g/mol). The

product was then dried in an oven at 65 °C until constant weight and then

grounded.

Characterization. The DMA conversion was measured by using Gas

Chromatography (GC). The samples were injected on a Hewlett Packard 5890

GC with an Elite-Wax ETR column with pentadecane as an internal standard.

The total molecular weight (Mn,tot) is calculated using the DMA conversion

(monomer-initiator ratio multiply by the conversion). The span molecular

weight (Mn,SPAN) is calculated using the Mn,tot and is defined as two times the

molecular weight of one arm plus the molecular weight of the macro-initiator

(comb PDMA).

Gel permeation chromatography (GPC) analysis of one (entry 6, Mn,th

falls in the range of the calibration curve of the GPC while the Mn,th of the rest

of the entries are all higher than the range) of the water-soluble samples was

performed on a Agilent 1200 system with Polymer Standard Service (PSS)

columns (guard, 104 and 103 Å) with a 50 mM NaNO3 aqueous solution as

the eluent. The columns were operated at 40 °C with a flow-rate of 1.00

ml/min, and a refractive index (RI) detector (Agilent 1200) was used at 40

°C. The apparent molecular weights and dispersities were determined using a

polyacrylamide (PAM) based calibration with WinGPC software (PSS).

Carbon (13) nuclear magnetic resonance (13C-NMR) spectroscopy was

performed on a Varian Mercury Plus 500 MHz spectrometer. For analysis D2O

was used as the solvent. The delay time was set at 2s and at least 10000

scans were performed (overnight). The polymer samples were swelled for 1

day and stirred for another day at room temperature. In order to obtain a

high signal to noise ratio, a high polymer concentration was used. The

hydrolysis degree was determined through the integration method reported

in literature.72

Page 32: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

209

Rheological characterization. The aqueous polymeric solutions were

prepared by swelling the polymers in water for one day and afterwards gently

stirring the solution for another day. Viscometric measurements were

performed on a HAAKE Mars III (ThermoScientific) rheometer, equipped with

a cone-and-plate geometry (diameter 60 mm, angle 2°). Flow curves were

measured by increasing the shear stress by regular steps and waiting for

equilibrium at each step. The shear rate ( ) was varied between 0.1 – 1750

s-1. Dynamic measurements were performed with frequencies ranging

between 0.04 – 100 rad/s (i.e., 6.37·10-3 – 15.92 Hz). It must be noted that

all the dynamic measurements were preceded by an oscillation stress sweep

to identify the linear viscoelastic response of each sample and to ensure that

the dynamic measurements were conducted in the linear response region of

the samples.

Alkaline hydrolysis. Stock solutions of the different polymers were

prepared by swelling the polymers for a day in the alkali-salt mixtures and

gently stirring for another day. The polymer concentration was set at 5000

ppm. NaHCO3 was used as the alkali agent and the concentration was fixed at

3000 ppm. One solution further contained 5000 ppm NaCl and the other one

contained 4925 ppm NaCl and 75 ppm CaCl2. The solutions were divided into

8 different vials (sealed) and placed in an oven at 70 °C. At set time intervals

a sample vial was removed from the oven and cooled to room temperature.

The viscosity function of the sample was then recorded. Both solutions were

evaluated for a total of 192 hours. The viscosity retention was evaluated for

the samples using equation 8.A.3:

(8.A.3)

where = the solution viscosity (measured at = 10 s-1) of the virgin

polymer samples and = the solution viscosity (measured at = 10 s-1) of

the polymer sample treated for the specified number of days.

Four other solutions were prepared in order to evaluate for longer

periods. Two sets of comparison were performed. In the first one, the

polymer concentration (in terms of monomeric moles) was set equal between

the different polymers and the other one the solution viscosity (at = 10 s-1)

was kept equal. Again NaHCO3 was used as the alkali agent and the

concentration was fixed at 3000 ppm. One solution further contained 10000

ppm NaCl and the other one contained 9850 ppm NaCl and 150 ppm CaCl2.

The solutions were divided into 8 different vials (sealed) and placed in an

oven at 70 °C. At set time intervals a sample vial was removed from the

Page 33: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

210

oven and cooled to room temperature. The viscosity function of the sample

was then recorded. Both solutions were evaluated for more than 63 days.

Page 34: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Chapter 8

211

Appendix 8.B

8.B.1. Experimental section

Chemicals. Cyclo octane (≥ 99,5 %), guar gum, methyl cellulose,

xanthan gum and sodium carboxy methyl cellulose were purchased from

Sigma Aldrich. The crude oil is a medium oil (API gravity equals 27.8) and

originates from the Berkel oil field in the southwest of the Netherlands. The

viscosity of the oil is 71 mPa.s at 20 °C. The branched non-ionic water

soluble (co)polymers used in the flow cell were previously synthesized using

atomic transfer radical polymerization (Chapters 6, and section 8.4).

Solution preparation. The polymeric solutions were prepared by

swelling the polymers for at least 12 hours in demineralized water and

subsequently stirred for another 12 hours.

Rheological characterization. Viscometric measurements were

performed on a HAAKE Mars III (ThermoScientific) rheometer, equipped with

a cone-and-plate geometry (diameter 60 mm, angle 2°). Flow curves were

measured by increasing the shear stress by regular steps and waiting for

equilibrium at each step. The shear rate was varied between 0.1 – 1750 s-1.

Dynamic measurements were performed with frequencies ranging between

0.04 – 100 rad/s (i.e. 6.37·10-3 – 15.92 Hz). It must be noted that all the

dynamic measurements were preceded by an oscillation stress sweep to

identify the linear viscoelastic response of each sample and to ensure that

the dynamic measurements were conducted in the linear response region of

the samples.

Flow-cell experiments. A schematic presentation of the flow-cell (with

the dimensions) is given in Figure 7.2. The flow cell has been adapted from

the original ones presented in literature73 to resemble dead ends (Figure 1.4)

that are present in oil reservoirs. The bottom part of the flow-cell is made out

of aluminum while the cover is glass. The depth of the chamber (designated

as blue in Figure 7.2) is set at 0.5 mm. The chamber is first filled with oil and

afterwards flooded with water or polymer solutions. For the branched PDMA

crude oil was used, and for the copolymers of AM and NIPAM a 1-2

(volume%) mixture of cyclo-octane and crude oil was used ( = 17 mPa.s).

The linear velocity was set at 1 foot per day. Each flood (either water or

polymer) was continued for at 24 hours at room temperature (RT). The oil

recovery out of the different cells was visually determined by taking high

definition pictures (before and after the floods). Analysis (pixel count) of the

image using Adobe allows the calculation of the amount of oil left behind in

the flow-cell.

Page 35: University of Groningen Synthesis and evaluation of novel linear … · solution viscosity) towards higher temperatures (T > 50 °C). The dependency of the solution viscosity on the

Outlook on the application of branched (co)polymers in EOR

212

This page intentionally left blank


Recommended