Sie sind auf Seite 1von 18

Mechanisms of Formation Damage:

Solids Plugging

Fig 1. Illite

Clay Swelling: Formation damage that occurs when water-based filtrates from drilling, completion, workover
or stimulation fluids enter the formation altering the ionic environment of clays via ion exchanges, changes in
pH, and/or changes in salinity, thus leading to a reduction in porosity and permeability. Figures 1A-1E show
SEM photographs of several different clays:

Figure 2: Smectite
Fig. 3 Kaolinite

Fig. 4 Chlorite
Fig. 5 Smectite-Illite

Smectite, for example, possesses negative charges on the faces of the clay crystal, while the edges are positively
charged. The density of negative charges on the clay structure is determined in terms of the cation exchange
capacity (CEC); which is the amount of positively charged ions (cations) that the clay structure can
accommodate on its negatively charged exterior. Thus, CEC values are a measure of the clay's propensity to
swell under aqueous conditions. Table 1 CEC ranges of several clays.

Clay Type CEC Range (Meq/100g)

Smectite 80 - 150

Kaolinite 3 - 15

Illite 10 - 40

Chlorite 10 - 40

Table 1: Cation Exchange Capacity Ranges of Several Clays 1

A common laboratory method for measuring CEC is through multiple salinity tests, a technique used for the
determination of the electrical properties of shale containing core samples. In this test, the sample is flushed
with brines of different salinities, and the conductivity determined after each flush. A plot of the conductivity of
the sample versus the conductivity of the brine gives the excess conductivity caused by clays and other surface
conductors. Then, using a suitable model (e.g., Waxman-Smits, dual water, SGS) it is possible to determine the
intrinsic formation factor and porosity exponent, and the cation-exchange capacity.2 Another source of clay
disturbance is usually associated with changes to the wetting phase (often native formation water). In the case of
smectite and mixed-layer clays (primarily smectite-illite), a change in size due to swelling or water retention
enhances their probability of getting dislodged and migrate with the mobile wetting phase. This phenomenon is
referred to as swelling-induced clay migration.3

Fines Migration: Formation damage which arises when the drawdown forces during flowback or production
exceed the cohesive forces between fines and the rock fabric. This in turn causes particles suspended in the
produced fluid to bridge the pore throats near the wellbore, reducing well productivity. Fines can include
different materials such as clays (phyllosilicates smaller than 4 microns) and silts (silicates or aluminosilicates
with sizes ranging from 4 to 64 microns).4 Fines migration can be exacerbated by the use of incompatible fluid
treatments. Commercial products have been developed to minimize the potential for fines migration (e.g. resin
consolidation, tackifiers, and covalent bonding of polymers).

Ultra-thin tackifying agents (UTTA), like Halliburton's SandTrap®, have been developed to stabilize fines in
high-rate producing or injection wells. These systems can be applied during initial fracturing or gravel-packing
operations, as a remedial treatment, or as a follow up to fracturing or acidizing treatments. Schlumberger's K300
is an example of technology based on the polymerization of resin. All of the developed products address the
issue of fines migration and have similar solutions. They all involve the application of some form of coating to
adhere fines to the mineral surfaces. Advantages of resin consolidation are that it is suitable for through-tubing
applications, applicable in small diameter casing, and that it can be applied in abnormal pressure well.5

Problems arise because resin consolidation involves multistage processes in which several fluids must be
uniformly applied sequentially into a perforated interval and frequently are highly toxic and relatively
expensive. Moreover, resin consolidation can significantly reduce the permeability to oil by changing the
wettability of the rock and by occluding the pore space with resin.

Sand Production: Formation damage that results from the production of sand and its subsequent movement
into pore throats and ⁄ or proppant-packs, causing plugging and productivity impairment. Key factors
influencing sand production are:

 Degree of formation consolidation which depends on cementation of sand grains around the perforation
tunnel, the geological age, and depositional environment.
 Reduction in pore pressure throughout the life of the well which results in an increasing amount of stress
on the formation sand, causing it to break loose from the matrix and get crushed, thus creating movable
fines that are produced along with the wellbore fluids.
 Production rate of reservoir fluids which creates a pressure gradient and frictional drag forces that
exceed the formation strength. Thus, there is a critical flow rate below which these forces will not
exceed the formation strength.
 Reservoir fluid viscosity which plays a vital role in the case of heavy oil reservoirs with low-gravity,
high-viscosity oils even at low production rates.
 Increase in water-cut influences sand production twofold. On one hand it decreases the relative
permeability of oil over the time after production, thus increasing the pressure differential and induced
stresses required to produce the well at the same rate, yielding sand production. On the other hand it
increases the likelihood of water-wet particles to move along with the aqueous (wetting) phase.

Sand production is detrimental to productivity over the life of the well. Some of the issues seen with sand
production are:

Plugging of perforations, reducing production efficiency.

 Erosion in surface and downhole equipment when the velocity of sand is high, increasing the need for
workover treatments.
 Collapse of formation may take place due to void formation around perforation tunnel over the time as
sand is being produced, decreasing permeability and increasing pressure drop.

Sand control can be achieved through various means including reducing drag forces (i.e. lower production
rates), mechanically bridging sand (e.g., gravel packs), and resin consolidation. An example of resin
consolidation is the silanol resin consolidation system. This sand control technology is a resin system consisting
of aromatic polyester amide and tri-alkoxy organosilane. The tri-alkoxy organosilane acts as a coupling agent
between the reservoir sand grains and aromatic polyesteramide which acts as the load bearing resin due to the
pore pressure gradient and overburden stresses. It is applicable at high pressures and temperatures, from about
50°F to 450°F. When in contact with formation water, the chemicals react to hydrolyze it at the specific sites to
form silanol glue which bonds the sand grains together, forming a strong bond. Additional information can be
found in SPE paper 120472.

Perforating Charge Debris: Formation damage caused by perforating is one of the highest risks in well
completions. As shown in Figure 26 (Published by Schlumberger; Used courtesy of Schlumberger;
Permission obtained Sept. 9, 2009), common types of damage that can occur inside the perforation tunnel are
fractured and compacted zones, perforation gun debris, and the reaction of perforating charge liner materials
(e.g. zinc) with high density brines upon detonation, seen in Figure 3.

When sprayed into clear completion brines at a high detonation temperature, high surface area particles become
activated and then react with the aqueous phase to form metal oxides, metal hydroxides and hydrogen gas. For
instance, when calcium chloride completion brines are used along with perforating charge cases containing zinc
alloy materials, a number of chemical reactions may take place, resulting in the formation of cementing
materials that can significantly block pore throats (SPE paper 58758). Equations 1A-1D demonstrates the
sequence of chemical reactions leading to the formation of cement type materials. Moreover, the chemical
nature of the reaction products suggests that typical scale inhibitors might function to reduce interparticle
associations and minimize the cementing or agglomeration process.7

Zn° + H2O → H2(g) + ZnO(ppt)


Equation 1A: Zinc Oxide Precipitate Formation Reaction
Zn° + 2H2O → H2(g) + Zn(OH)2(ppt)
Equation 1B: Zinc Hydroxide Precipitate Formation Reaction

Zn° + CaCl2 + 2H2O → ZnCl2 + Ca (OH) 2(ppt) + H2 (g)


Equation 1C: Zinc Chloride & Calcium Hydroxide Precipitate Reaction

xZn (OH)2 + yZnCl2 + zH2O → 2Zn(x+y)(OH)xCly(H20)z(ppt)


Equation 1D: Complex Zinc Hydroxy Chloride Precipitate Formation Reaction

High-temperature ⁄ high pressure (HTHP) wells are particularly susceptible to this source of damage. A post
perforating acid treatment can be performed in order to revert some of the damage; however, as formation
temperatures increase, metal corrosion and acid sensitivity of the formation become problematic. At higher
temperatures, organic acids are frequently used, but many of them do not have the acid strength or the capability
to dissolve zinc or zinc salts. The long-chained organic acid HTO has been shown to dissolve zinc and
perforating gun debris. The solubilities of zinc metal and gun debris at 250° F (121° C) are shown in table 2. It
is estimated that typical weights of debris can range from 0.2 lb/ft (1.4 kg/m) in low debris carriers to 1.4 lb⁄ft
(13 kg⁄m) in steel carriers at a 12 shot⁄ft density. At the higher temperatures, above 250° F (121° C), a savings
of up to 20% on acid volume can be realized based on the increased dissolving power of a long chained organic
acid (e.g. HTO). Table 2 shows the solubilities of zinc and gun debris in different acids at different
temperatures.

10% Formic Acid 10% Acetic Acid 10% HTO Acid

Zn metal @ 21° C 0.10 lb⁄gal 0.02 lb⁄gal 0.09 lb⁄gal

Zn metal @ 121° C 0.28 lb⁄gal 0.24 lb⁄gal 0.34 lb⁄gal

Gun Debris @ 21° C 0.23 lb⁄gal 0.27 lb⁄gal 0.17 lb⁄gal

Gun Debris @ 121° C 0.28 lb⁄gal 0.27 lb⁄gal 0.28 lb⁄gal

Table 2: Zinc & Gun Debris Solubilities in Various Acids8

Particle Precipitation: Formation damage caused by the formation of an insoluble material in a fluid. Particle
precipitates can be classified as organic, inorganic, or organometallic.

Inorganic:

Calcite: Calcium carbonate (CaCO3) scale, the most common inorganic scale, precipitates as pressure is
reduced and CO2 is given off from the formation water and calcium scale is deposited. The production of scale
produces a further drop in reservoir pressure, causing more scale to be formed. The deposition takes place
through the following reaction (Equation 2):
Ca2+ + 2HCO3 → CaCO3 (s) + CO2 (g) + H2O

Equation 2: Calcium Carbonate Scale Formation Reaction

Induced scaling also occurs by mixing of formation brine with extraneous incompatible fluids invading the
reservoir during drilling, cementing, completion, and workover operations. For the example above, any increase
of the dissolved calcium (Ca2+) cation concentration caused by these operations is compensated by calcium
carbonate (CaCO3) precipitation.9 Effective calcium carbonate scale removal can often be achieved through
acid treatments, as CaCO3 is highly soluble in acid. However, spent acid can contain high concentrations of
scale producing ions, often leading to short lived stimulation treatments as the calcium carbonate re-precipitates
around the near wellbore region. Also effective are chelating agents, but they can be expensive. Chelating
agents work by preventing the chelated Ca2+ cations from re-precipitating after treatment. In order to prevent
calcium carbonate scaling, inhibitors squeezes have been used. These treatments work by either adsorbing onto
the formation material, providing a prolonged treatment through desorption into production fluids, or through a
precipitation mechanism. The precipitation mechanism functions by precipitating a calcium salt into the pores
which dissolve over time during production, providing inhibition. This method might increase treatment life,
but also presents the possibility of inducing damage into the producing formation.10

Barite Scale: Barium sulfate (BaSO4) scale formation occurs when the concentration of barium sulfate exceeds
the saturation point, causing the excess BaSO4 to precipitate. The saturation point of an aqueous solution
dependent upon temperature, pressure, and solvent composition. Solubility of barium sulfate increases with
temperature, pressure, and salt content of the brine. Factors that commonly induce BaSO4 are lower
temperatures, brine dilution, pressure drops, and mixing of incompatible waters. The deposition takes place
through the following reaction (Equation3):

Ba2+(aq) + SO42-(aq) → BaSO4

Equation 3: Barium Sulfate Formation Reaction

Barium sulfate scale is especially difficult to remove through acid treatments due to the high cost of treatments.
However, EDTA and nitrilotriacetic acid (NTA) are two chemicals that can be used for removal. Mechanical
removal and coiled tubing operation are the only effective methods of BaSO4 scale removal. Laboratory test
should be performed to determine the inhibitor concentration needed to prevent barium sulfate scale formation
and to evaluate the effectiveness of the inhibitor as changes in temperature, pH, and salinity. Inhibitors
commonly used are phosphonates, phosphate esters, polyphosphonates, and polymeric species. Additional
treatments can include: squeeze treatments, continuous injection (upstream of known risk points, capillary
string injection), precipitation squeezes (where scale inhibitor precipitates and dissolves slowly over time into
the brine), solid inhibitors (placed in the rat hole, associated with proppant), scale inhibitors included in
hydraulic fluids, or gas lift deployed inhibitors.11

Anhydrite Scale: Calcium sulfate (CaSO4) scale deposition is largely dependent upon pressure changes. The
deposition takes place through the following reaction (Equation 4):
Ca2+(aq) + SO42-(aq) → CaSO4(s)

Equation 4: Calcium Sulfate Scale Formation Reaction

Temperature is also a factor, with higher temperatures lowering the anhydrite solubility and increasing scaling
tendency. In seawater injections, scale such as anhydrite will become more significant as seawater breakthrough
occurs. There are 3 available methods for chemical removal of anhydrite scales.12,13

 Inorganic converters, which modifies the scale into an acid soluble byproduct. This method will also
remove other acid soluble materials present.
 Organic converters, which converts the scale into a dispersion/sludge that is able to flow. This method
can include an acid treatment or not. The acid treatment will effectively remove the reaction products
because they are soluble in acid.
 Chelants, which work by complexing the Ca2+ ions. This method effectively reduces the ions capacity to
re-precipitate. Inhibition of anhydrite scale could involve polyphosphonates or polyorganic acid salt
compounds.

Halite Scale: Salt scale that can be formed during production of high salinity (>200,000 ppm) formation brine
as seen in Figure 8. Halite formation may also occur during the evaporation of water into the gas phase.
Halite scale is normally easily removed with periodic fresh or low salinity water flushes. Removal can also be
achieved with continuous dilution of the fluid stream with water upstream of where deposition occurs. 14
Depending on the rate of the salt deposition and the availability of fresh water, such flushes could become an
expensive removal method. An example of a salt inhibitor used is potassium hexacyanoferrate (HCF). HCF is a
well-known species which has been applied as an anti-caking agent in cooking, and as a drilling-fluid additive
for drilling through salt layers, where it both limits hole wash-out (because it also reduces the rate of salt
dissolution) and prevents salt from crystallizing from the returned fluid as it cools and becomes supersaturated
in salt.15

Iron Sulfide: Scale that can occur whenever sources of both iron and hydrogen sulfide are present. H2S can
result from the presence of sulfate reducing bacteria, thermal sulfate decomposition or introduction to a well
through gas lift operations. Iron sulfides are able to enhance the corrosion process, decrease productivity, and
negatively affect oil-water separation activities. Iron sulfide exists in numerous crystalline forms with numerous
acid solubilities. The FeS species responds well to HCl treatment, but the longer the contact time between FeS
and H2S the more likely that the scale will become richer in sulfur. While FeS may be effectively removed with
acid, FeS2 is not. Since iron sulfide is normally oil-wet, scale removal is impeded. To correct this, adding
surfactants and water-wetting agents is important. Acid treatments should also have a corrosion inhibitor, an
iron control agent, and a hydrogen sulfide scavenger. Toxic H2S is produced by the following reaction
(Equation 5) between FeS and HCl:

FeS(s) + 2HCl (aq) → FeCl2 (aq) + H2S (g)

Equation 5: Hydrogen Sulfide Gas Formation Reaction


FeS will also precipitate as H2S continues to react with any ferrous iron present at pH > 1.9. If ferric ion is
present elemental sulfur can precipitate, which is insoluble in HCl and needs expensive organic solvents to
remove.16 Understanding the source of iron and sulfide is key to preventing iron sulfide scaling. Iron can be
present in the formation water or supplied by tubing corrosion. If the iron is supplied by tubing corrosion,
protecting the metallurgy could reduce the potential for iron sulfide scale. If the iron is present in the formation
water, the course of action should be to limit the amount of H2S through biocides, injection water sulfate ion
minimization, or injection of nitrates. One chemical treatment option is tetrakis hydroxymethyl phosphonium
(THPS), used to dissolve or chelate iron sulfide once it is formed.17

Organometallic:

Naphthenates: Formation damage caused by fluctuations in the reservoir water pH, resulting in the formation
of organic scales, carbonate deposits, and the stabilization of emulsions. Reservoir water is naturally saturated
with CO2 in equilibrium with bicarbonate anion (HCO3-) as shown in the following reaction (Equation 6):

CO2 + 2H2O → HCO3- + H3O+

Equation 6: Reservoir Water Equilibrium

Fluids injected into the well for various procedures can alter the temperature, pressure, and composition of the
fluids in the near wellbore region. Precipitation can occur during production by a chemical reaction of two or
more ions in solution or by changing the temperature⁄pressure of a saturated solution which causes a drop in
solubility. Scale can also precipitate due to the mixing of two incompatible fluids, and with the release of CO 2
brought on by a pressure reduction.

These pressure drops are accompanied by an increase in pH and oftentimes, the formation of mixed carbonate
and naphthenate deposits inside tubing or surface installations, as well as the creation of stable emulsions due to
the surface-active naphthenate group RCOO-. Naphthenic acids, R-COOH, are often present in crude oils and the
hydrophilic nature of the carboxylic acid group means that they congregate at the oil-water surface.18 Examples of their structures
can be seen in Figure 11. Oil and formation water composition is very important in the formation of naphthenates. These variables
are naphthenic acid concentration and composition, formation water cations, bicarbonates and pH. Crude oils that present the biggest
complications are ones with high total acid number, TAN, and high concentrations of naphthenic acid.

Naphthenate problems can be exacerbated by the presence of solids such as formation sand and fines, waxes, and other types of scale.
The stability of emulsions containing naphthenic acids has been shown to be a function of pH, asphaltene⁄resin ratios, naphthenic acid
types, and cation content of the aqueous phase. Sodium rich emulsions lead to less separated water volume over time, showing the
stability of the oil-water emulsion. Calcium rich solutions lead to less stable emulsions, possibly due to excess ionic strength in
solution. Sarac and Civan19 determined through experimentation that the critical minimum initial brine pH required for the onset of
naphthenate precipitation to be 5.91. As pressure drops occur during production, degassing of CO 2 takes place, raising the pH of the
formation brine and promoting the dissociation of naphthenic acids as shown in Equation 7.
R-COOH + OH- → R-COO- + H2O

Equation 7: Naphthenic Acid Dissociation

The naphthenate ion is very reactive and tends to complex with Na+ and Ca2+ cations to form sodium and
calcium naphthenate scales as shown in Equation 8 and Equation 9. Naphthenate deposits normally collect in oil
⁄ water separators but can deposit in tubing and pipelines as well.

2R-COO- + Ca2+ → (R-COO)2Ca

Equation 8: Calcium Naphthenate Formation

R-COO- + Na+ → R-COONa

Equation 9: Sodium Naphthenate Formation

Due to its high molecular weight, calcium naphthenate is less soluble than sodium naphthenate in water. This is
important because when calcium carbonate and calcium naphthenate form together, the carbonate will decrease
the formation of naphthenate. This is due to the reduction in available calcium cations for reactions with the
naphthenate anions. When evaluating the stability of emulsion and the amount of naphthenate deposits during
processing of acidic crude it is important to take into account the following criteria:

1. Water pH value at process conditions as well as the level of bicarbonate and calcium content at reservoir
conditions.
2. Total Acid Number (TAN) of crude oil.20 TAN is the amount of any acid contained in an oil sample.
While the test is unable to determine specific types of acids, it is useful in determining if a sample of oil
will be corrosive or not. The threshold for corrosive oils is 0.5 mg KOH⁄g oil. Acidizing with HCl and ⁄
or acetic acid is often used to remove naphthenate deposits. An example of naphthenate deposits can be
seen in Figure 12. Additional information on naphthenate formation, prevention and mitigation can be
found in SPE papers 93407, 80395, 112434, and 68307.

Organic:

Asphaltenes & Parrafins: Formation damage resulting from organic deposit which hamper the production of
crude oil. Paraffins are alkanes of relatively high MW (C18 to C70), which can be either straight-chained or
branched. They have specific solubilities and melting points. Because these hydrocarbons have satisfied valence
electron configurations, they are almost completely inert to chemical reactions and, as a result, immune to
attack by bases and acids.

Paraffin waxes are soluble in most liquid petroleum fractions, and their solubility normally decreases as MW
increases. Hence, they are soluble in both straight-chain and aromatic petroleum derivatives. They are deposited
as solids when the temperature drops below the cloud point for the particular crude oil. SARA analysis can be
performed to determine the different constituents present in oil. SARA is a method for characterization of oils
based on fractionation, whereby a heavy oil sample is separated into smaller quantities or fractions, with each
fraction having a different composition.

Fractionation is based on the solubility of hydrocarbon components in various solvents used in the test. Each
fraction consists of a solubility class containing a range of different molecular-weight species. In this method,
the oil is fractionated to four solubility classes, referred to collectively as SARA: Saturates-Aromatics- Resins-
Asphaltenes.

Saturates are generally paraffins, while aromatics, resins, and asphaltenes form a continuum of molecules with
increasing molecular weight, aromaticity, and heteroatom contents. Products like Halliburton's Parachek® 160,
a polymeric paraffin inhibitor, alter the paraffin structure, decreasing its tendency to precipitate. 21 Common
solvents used for paraffin removal include condensate, kerosene, and diesel (straight-chain hydrocarbons).
Asphaltenes on the other hand are black, polycyclic aromatic, complex compounds, seen in Figure 5.
Published by Experimental Soft Condensed Matter Group ( Harvard); Permission to use obtained Aug. 3,
2009. Generally, they are spherical, 30Å to 65Å in diameter, with MW of 10,000 to 100,000.

These molecules are held in suspension by surrounding asphaltic resins (maltenes). Asphaltenes polar
properties result from the presence of oxygen, sulfur, nitrogen and various metals in their structures. Figure 7
shows the blockage that can occur resulting in severe damage. Published by London Center for
Nanotechnology; Permission to use obtained Sept. 9, 2009. Deposition occurs not only by temperature ⁄
pressure reductions, but also by destabilizing factors which act on the resins such as contact with acid, CO 2, or
aliphatic solvents, that act on the micellar colloidal suspension to strip away the maltene and resin from the
micelles. Removal is possible, but selecting the appropriate method is crucial and can be accomplished by field
tests.

While condensate, kerosene, and diesel are commonly used to dissolve paraffin, they should not be used when
attempting to remove asphaltenes. These non-aromatic hydrocarbons, if used, can cause further precipitation of
the asphaltenes as the maltene stabilizers are disturbed. Instead, aromatic chemicals such as xylene can be used.
Their power can be enhanced by almost ten times with the addition of approximately 5% by volume of a
specific primary or secondary amine, such as Halliburton's highly polar organic Targon® II. 22 Closely
monitored, due to low flash points, moderate heating will hasten the removal process. New solvents that are
non-toxic, biodegradable, and work similarly are available, such as Dowell Schlumberger's PARAN ECO®.
Another asphaltene solvent is Tretolite's Parid® PD-72, which is a mixture of toluene and a surfactant. While
continuous or batch pumping methods are employed, the batch method is recommended with the solvent left in
contact with the asphaltenes for up to 24 hours. Common methods for organics scale removal are mechanical,
solvents, heat, and dispersants.

Hydrates: Hydrates are solid, white, crystalline substances, with cellular structures, formed as a result of water
vapors and gaseous hydrocarbons interaction in the presence of water and under high pressure and low
temperature conditions (T > 32° F) as shown in Figure 9. There are three common types of gas hydrate
inhibitors; thermodynamic, kinetic, and crystal size modifiers. Thermodynamic inhibitors (e.g., inorganic and
organic salts, glycerol or low molecular weight glycol, combination of salt and glycol) work upon injection by
preventing the formation of hydrogen bonds or destroying them. Kinetic inhibitors (e.g., polymers and
surfactants) work as a slow reaction to delay nucleation or slow the crystal growth rate. Crystal size modifiers,
also known as crystal habit modifiers, do not prevent hydrate formation. Instead, they act as anti-agglomerates
to ensure that hydrates form a pumpable slush so that fluid flow is maintained.

Bacterial Growth

Formation damage that is caused by the introduction of bacteria while aqueous phase fluids are utilized and
improper bacteriological control is maintained. Bacterially induced formation damage is a particularly insidious
type of formation damage in that the apparent harmful effects of the introduction of the bacterial agents are
usually not noticed until well construction materials fail catastrophically. Bacteria which cause formation
damage can be classified into two types, aerobic and anaerobic. Aerobic bacteria require a constant source of
oxygen to survive and are mainly problematic with long term water injection operations. Anaerobic bacteria do
not require oxygen and tend to be more widespread and problematic. Both types present issues with plugging,
corrosion and toxicity.23 Corrosion is caused by anaerobic bacteria, sulfate reducers, which digest sulfate in
water to produce corrosive hydrogen sulfide. The resulting iron sulfide corrosion product, particularly in
combination with small amounts of oil, can significantly plug water treatment and injection facilities. Slug
treatments with bactericide are usually effective in controlling the anaerobic sulfate reducers.24 Other
alternatives (still in early development stages) rely on the use of phage cocktails to target specific bacteria
within the reservoir and/or in pipe lines.25

Polymer Plugging

Formation damage caused by the addition of polymers typically used to provide clay stability and ⁄ or control
fluid losses during drill-in and completion operations. Chemical fluid loss control (FLC) materials can be
grouped into two categories, solids laden and solids free. Sized salts, calcium carbonate, and organosoluble
resins are three types of solids typically used as FLC materials. Solids free FLC pills, on the other hand, may
consist of linear gels (e.g., Liqui-Vis® and Bromi-Vis®), crosslinked gels, as seen in Figure 13 (e.g., K-
Max™, Max Seal™, TekPlug™, and Protectozone™), and the more recently developed viscoelastic surfactant
gels (e.g., ClearFrac™).

 Solids Laden FLC: The most widely used solids laden FLC materials consist of sized CaCO3 particles
suspended in a polymer matrix. Computer software is used to determine the optimum CaCO3 loading
and particle size distribution to form a seal against the formation rock and minimize fluid losses. The pill
can be either bullheaded or spotted with a coil tubing unit. The pill is usually removed with an acid
treatment (e.g., HCl or HCl⁄acetic). Other clean-up alternatives include breaker systems (internal or
external) based on polymer-specific enzymes, and ⁄ or chelating agents. An alternative to the calcium
carbonate⁄polymer system, particularly for injector wells, is the salt⁄polymer system which consists of
ground sodium chloride in saturated NaCl brine ⁄ polymer matrix. Since sodium chloride is readily
soluble in water, produced water or unsaturated brine treatments will afford removal of residual salt
solids. Depending on the polymer uploading in the FLC pill, an acid treatment may also be required.
Santrol's Collagen or synthetic polymer balls are also readily dissolved in the presence of unsaturated
brines, and thus constitute another alternative for water-soluble FLC materials, particularly for injector
wells.26 Oil soluble resins such as benzoic flakes, may be used as FLC materials in low to moderate
temperature producer wells. Resins are typically added to completion brines and delivered to the
formation where they plate out onto rock surfaces. After the completion operation is finished, produced
oil or condensate flowing over the resin gradually dissolves it. As with the introduction of any foreign
fluid into the formation, there are advantages and disadvantages to be evaluated before using a FLC
material. Table 3 highlights the strengths, limitations, and costs of the three general types of bridging
solids used for FLC.

Particulate Strengths Limitations Cost

CaCO3 ⁄ Inexpensive Can plug perforation tunnels Relatively


Polymer Relatively easy to mix Highly damaging if particles are not sized low
Good for bridging on gravel and properly
frac packs Polymer damage may require remedial
Used when lower densities are treatments
required May cause tools to stick
Degradation depends upon contact with
acid for removal

NaCl ⁄ Relatively easy to mix Can plug perforation tunnels Low


Polymer Good for bridging on gravel and Highly damaging if particles are not sized
frac packs properly
Water soluble particles Polymer damage may require remedial
treatments
May cause tools to stick

Oil Soluble Easy to mix May not stop losses completely Relatively
Resin Good for bridging on gravel and Not useful above melting point low
frac packs Only recommended in oil wells
Water soluble particles

Table 3: Bridging Particulate Systems Strengths, Limitations, & Costs

Solids Free FLC: Linear gels can be prepared from a range of polymer systems, most typically natural
polymers or natural polymers that have been modified to achieve the necessary purity and solution properties.
The chemical structure of common polymers used as FLC materials and viscosifiers are shown in figures 14A-
14F. Cellulose, Figure 14A, is a natural structural component of wood and in cotton it exists in nearly pure
form. However, cellulose is insoluble in water and brines. To make this polymer soluble, it is usually
derivatized to hydroxyethyl cellulose (HEC), Figure 14B, which is relatively easy to disperse and hydrate in
most brine solutions. Xanthan polymer, Figure 14C, is a more complex system, with excellent solids
transport and suspending characteristics, particularly at low shear rates. Note that xanthan has pendant carboxyl
groups, which can bind to contaminant ions to produce difficult-to-break gels. Additionally, xanthan polymer
systems are incompatible with many biocides and clay stabilizers. Linear polymer gels are useful in operations
involving low overbalance, low temperature, low permeability, and short interval length. More recently,
viscoelastic surfactants (VES) have been developed as an alternative to polymer-based FLC materials. VES
impart viscosity to the fluid by forming a 3D network of rod-like micellar structures, show in Figure 14E.
The advantage of viscoelastic surfactants is that they need no added breakers to reduce viscosity after use.
Viscosity is degraded when contact is made with oil, whereby the micelles are disrupted and the viscosifying
network is destroyed. Another benefit of the particles is the formation of a pseudo filter cake of viscous VES
fluid that greatly decreases fluid loss rates and improves overall efficiency of the fluid. Table 4 lists the
strengths, limitations, and cost of some available linear gel systems and VES. New technology has led to the
introduction of nanometer-scale particles, which interact with VES micelles, through chemisorptions and
surface charge attraction, to form a worm-like micellular structure as seen in Figure 14F (Published by
Schlumberger - Used courtesy of Schlumberger. Permission obtained September 9, 2009). Such interactions
stabilize fluid viscosity at high temperatures. As internal breakers are activated to break the micelles, the fluid
drastically loses its viscosity and the pseudo filter cake dissolves into nanometer-sized particles. Since these
particles are so small, they are easily carried back to the surface, thus minimizing formation damage potential.
SPE paper 107728 provides more information on the new VES system.

Polymer Strengths Limitations Cost

Hydroxyethyl Cellulose Relatively easy to mix Highly damaging if not pumped Relatively
(HEC) Readily available correctly inexpensive
Non-toxic, non-corrosive Does not stop losses completely
Acid soluble Requires multiple treatments
Internally or externally Requires shearing and filtering
broken

Xanthan (XC) Relatively easy to mix Highly damaging if not pumped Relatively
Readily available correctly inexpensive
Non-toxic, non-corrosive Does not stop losses completely
Good low-shear viscosity Requires multiple treatments
Requires filtering
Not readily removed with acid

Succinoglycan Easy to mix Does not stop losses completely Moderately


Non-toxic, non-corrosive Requires multiple treatments expensive
Good low-shear viscosity Requires filtering
Temperature limitations

Viscoelastic Surfactant Easy to mix Requires hydrocarbon contact to Relatively


(VES) No solids break expensive
Good low-shear viscosity Temperature limitations

Table 4: Linear Gel Systems (Polymers) Strengths, Limitations, & Cost

Figure 15 shows a comparison of rheological properties of HEC and xanthan (XC) polymer. As seen in this
figure, XC polymer displays a higher low shear rate viscosity (LSRV) compared to HEC, which translates in
better solids carrying capacity at low flow rates, and under static (no flow) conditions. On the other hand, HEC
is easier to remove via the use of internal breakers and ⁄ or remedial acid treatments. To achieve higher
viscosities, crosslinked polymer systems are typically employed. Crosslinked gels are typically compromised of
derivatized HEC and are useful at temperatures up to 300° F. They are available as pre-blended products or can
be prepared on the rig. Crosslinking agents for HEC include zirconium and lanthanide salts. Crosslinked gel
particulates, in the form of premanufactured slurries, are oftentimes useful when more complete fluid loss
control is needed. Max Seal™ slurries, for example, can effectively plug off formations under relatively high
overbalanced pressures. It is comprised of crosslinked HEC gel, which has been chopped up to form
inhomogeneous, lumpy, flowing slurry. Table 5 lists some crosslinked gel systems, along with their strengths,
limitations and relative cost.

Crosslinked Strengths Limitations Cost

Crosslinkable HEC  Stops losses completely  Internal breakers used for ≤ Moderately expensive
 Easily removed with breakers 48 hrs
(fluid begins to degrade with
 Can withstand high overbalanced
time)
pressures
 Must be mixed on location
 Can run tools through it

Pre-Prepared  Stops losses completely  No internal breakers can be Moderately expensive


Crosslinked  No on-site mixing required added to high
Gel Particles (must be broken externally)
 Can withstand high overbalanced
pressures  Bulky to ship and store
 Can run tools through it

Zinc Bromide  Stops losses completely  Must be mixed on site High


Crosslinked Gel  Only pumpable system available  Must be broken externally
for ZnBr2

Table 5: Cross linked Gel Systems (Polymers) Strengths, Limitations, & Cost

Another potential source of formation damage from polymer plugging is that related to clay stabilization via the
use of polymers. The polymers include cationic inorganic polymers (CIP) and cationic organic polymers (COP).
CIP such as hydroxyl aluminum and zirconium oxychloride provide marginally permanent clay stabilization.
They offer resistance to cation exchange, but their application is limited to noncarbonate-containing sandstones.
Cationic organic polymers are effective in providing permanent stabilization to clays, especially smectite, and
for controlling fines and sand in sandstone and carbonates formations. Permanent protection is provided by the
availability of multiple cationic sites of attachment, but their application is limited to low concentrations. COP
are applicable in acidizing and fracturing, but their effectiveness is lowered in gelled-water solutions used for
hydraulic fracturing and gravel-packing due to gel competition for adsorption on clay surfaces.27 They are often
the cause of formation damage via polymer plugging because of their high molecular weight and long chains
that have molecular sizes similar to pore throats in porous rock. Figure 16 shows a scanning electron
micrograph detailing solids and polymer plugging of pore throats.
Altered Wettability

Formation damage in which the formation wettability is modified, generating a change in relative permeability
to oil, gas and ⁄ or water that eventually affects well productivity. In particular, surfactants and other additives in
drilling fluids, especially oil-base mud, can change a naturally water-wet formation to an oil-wet formation with
consequent production impairment caused by reduction of relative permeability to oil and/or gas. Brine salinity
and pH are another important factor related to wettability because they strongly affect the surface charge density
on the formation rock and fluid interfaces, which in turn can affect surfactants adsorption. Figure 17 shows
an SEM of water droplets on both kaolinite and quartz, illustrating the contrasting wetting characteristics of
different mineral surfaces. Published by University of the West of Scotland; "No further reproduction, please
use for this project webpage only." Permission to use obtained Aug. 5, 2009. Relative permeability modifiers,
such as Halliburton's WaterWeb® and BJ Services' AquaCon™, are hydrophilic polymers designed to reduce
the effective permeability to water, while increasing (or maintaining) the relative permeability, to gas and ⁄ or
oil. They do not typically require special placement techniques.

Water Block

Formation damage that occurs when large quantities of water and/or brine are lost to the formation, thus
increasing water saturation and decreasing the relative permeability to oil and/or gas. Partially pressure-depleted
reservoirs are particularly sensitive to this type of damage. Water blocking can be prevented ⁄ minimized by
adding surface tension reducing agents (e.g. surfactants, alcohols, or microemulsions) to wellbore fluids to not
only lower surface and interfacial tension, but also to water-wet the formation, and prevent emulsions.28

Emulsions

Formation damage that is a mixture of two, or more, immiscible liquids in which the liquids are stabilized by
one or more emulsifying agents. If an emulsion block exists, well permeability as determined through injectivity
test will be much greater than permeability determined through production tests. Oilfield emulsion types consist
of water in oil (regular emulsions), oil in water (reverse emulsions) and complex emulsions. Most emulsions
break easily when the source of the mixing energy is removed. However, some natural and artificial stabilizing
agents, such as surfactants and small particle solids, keep fluids emulsified. Natural surfactants, created by
bacteria or during the oil generation process, can be found in many waters and crude oils, while artificial
surfactants are part of many drilling, completion or stimulation fluids. Among the most common solids that
stabilize emulsions are iron sulfide, paraffin, sand, silt, clay, asphalt, scale, and corrosion products. 29 More
information regarding emulsions can be found in SPE papers 105858, 100430, and 97886.
References:

1
Allen, Thomas O. and Roberts, Alan P. "Production Operations 2". OGCI, Tulsa, Oklahoma, 2000.
2
Allen, Thomas O. and Roberts, Alan P. "Production Operations 2". OGCI, Tulsa, Oklahoma, 2000.
3
Allen, Thomas O. and Roberts, Alan P. "Production Operations 2". OGCI, Tulsa, Oklahoma, 2000.
4
Schlumberger Oilfield Glossary
5
Allen, Thomas O. and Roberts, Alan P. "Production Operations 2". OGCI, Tulsa, Oklahoma, 2000.
6
Schlumberger: Perforation Damage
7
Javora, P.H.; Ali, S.A. and Miller, R. "Controlled Debris Perforating Systems: Prevention of an Unexpected Source of Formation Damage." SPE 58758,
2000.
8
McElfresh, P.M.; Gabrysch, A.D.; Van Sickle, E.; Myers, Jr. B. and Huang T. "A Novel Method of Preventing Perforation Damage in High-Temperature
Offshore Wells." SPE 86521, 2004.
9
Civan, Faruk. "Reservoir Formation Damage." 2nd Ed, Elsevier Inc., Oxford, UK, 2007.
10
Thane, C.G. "SMAD-Mafumeira Drilling, Completions and Production Formation Damage Review and Recommendations: Block 0, Angola." TM 2008-27,
2008.
11
Jordan, M.M; Collins, I.R.; Gyani, A. and G.M. Graham. "Coreflood Studies Examine New Technologies That Minimized Intervention Throughout Well Life
Cycle." SPE 74666, 2006.
12
Jewell, R.J. and B.R. Lasater. "New Products to Solve Scale Problems." SPE 3550, 1971.
13
O.J. Vetter. "Oilfield Scale – Can We Handle It?" SPE 5879, 1976.
14
Frigo, D.M.; Jackson, L.A.; Doran, S.M. and R.A. Trompert. "Chemical Inhibition of Halite Scaling in Topsides Equipment." SPE 60191, 2001.
15
Earl, S.L. "Use of Chemical Salt Precipitation Inhibitors to Maintain Supersaturated Salt Muds for Drilling Salt Formations." SPE 10097, 1981.
16
Nasr-El-Din, H.A. and A.A. Al-Taq. "Water Quality Requirement and Restoring the Injectivity of Waste Water Disposal Wells." SPE 68315, 1998.
17
Nasr-El-Din, H.A.; Roser, H.R. and Al-Jawfi, M. "Formation Damage Resulting from Biocide⁄Corrosion Squeeze Treatments." SPE 58803, 2000.
18
Dyer, S.J.; Graham. G.M. and Arnott, C. "Naphthenate Scale Formation – Examination of Molecular Controls in Idealised Systems." SPE 80395, 2003.
19
Sarac, S. and Civan, F. "Mechanisms, Parameters, and Modeling of Naphthenate Soap-Induced Formation Damage." SPE 112434, 2008.
20
Rousseau, G.; Zhou, H. and Hurtevent, C. "Calcium Carbonate and Naphthenate Mixed Scale in Deep-Offshore Fields." SPE 68307, 2001.
21
"Parachek® 160 Paraffin Inhibitor." Chemical Compliance. 08 Jan. 2008. Halliburton. 29 May 2009.
22
"Targon® II Asphaltene Solvent." Chemical Compliance. 08 Jan. 2008. Halliburton. 29 May 2009.
23
Bennio, D.B. et al. "Mechanisms of Formation Damage and Permeability Impairment Associated with the Drilling, Completion and Production of Low API
Gravity Oil Reservoirs." SPE 30320, 1995.
24
Allen, Thomas O. and Roberts, Alan P. "Production Operations 2." OGCI, Tulsa, Oklahoma, 2000.
25
Phage Biocontrol, Inc.
26
Santrol Oil & Gas Stimulation Products.
27
Civan, Faruk. "Reservoir Formation Damage". 2nd Ed, Elsevier Inc., Oxford, UK, 2007.
28
Allen, Thomas O. and Roberts, Alan P. "Production Operations 2." OGCI, Tulsa, Oklahoma, 2000.
29
Schlumberger Oilfield Glossary.

Last Update: 21 September 2009

Das könnte Ihnen auch gefallen