Handbook of Corrosion Engineering Episode 1 Part 5 docx

40 337 0
Handbook of Corrosion Engineering Episode 1 Part 5 docx

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

The PSI is calculated in a manner similar to the Ryznar stability index. Puckorius uses an equilibrium pH rather than the actual system pH to account for the buffering effects: PSI ϭ 2 (pH eq ) Ϫ pH s where pH eq ϭ 1.465 ϫ log 10 [Alkalinity] ϩ 4.54 [Alkalinity] ϭ [HCO 3 Ϫ ] ϩ 2[CO 3 2Ϫ ] ϩ [OH Ϫ ] Larson-Skold index. The Larson-Skold index describes the corrosivity of water toward mild steel. The index is based upon evaluation of in situ corrosion of mild steel lines transporting Great Lakes waters. The index is the ratio of equivalents per million (epm) of sulfate (SO 4 2Ϫ ) and chloride (Cl Ϫ ) to the epm of alkalinity in the form bicarbonate plus carbonate (HCO 3 Ϫ ϩ CO 3 2Ϫ ). Larson-Skold index ϭ As outlined in the original paper, the Larson-Skold index correlated closely to observed corrosion rates and to the type of attack in the Great Lakes water study. It should be noted that the waters studied in the development of the relationship were not deficient in alkalinity or buffering capacity and were capable of forming an inhibitory calcium carbonate film, if no interference was present. Extrapolation to other waters, such as those of low alkalinity or extreme alkalinity, goes beyond the range of the original data. The index has proved to be a useful tool in predicting the aggres- siveness of once-through cooling waters. It is particularly interesting because of the preponderance of waters with a composition similar to that of the Great Lakes waters and because of its usefulness as an indicator of aggressiveness in reviewing the applicability of corrosion inhibition treatment programs that rely on the natural alkalinity and film-forming capabilities of a cooling water. The Larson-Skold index might be interpreted by the following guidelines: Index Ͻ 0.8 Chlorides and sulfate probably will not inter- fere with natural film formation. 0.8 Ͻ index Ͻ 1.2 Chlorides and sulfates may interfere with nat- ural film formation. Higher than desired corro- sion rates might be anticipated. Index Ͼ 1.2 The tendency toward high corrosion rates of a local type should be expected as the index increases. epm Cl Ϫ ϩ epm SO 4 2Ϫ ᎏᎏᎏᎏ epm HCO 3 Ϫ ϩ epm CO 3 2Ϫ Environments 109 0765162_Ch02_Roberge 9/1/99 4:01 Page 109 Stiff-Davis index. The Stiff-Davis index attempts to overcome the shortcomings of the Langelier index with respect to waters with high total dissolved solids and the impact of “common ion” effects on the driving force for scale formation. Like the LSI, the Stiff-Davis index has its basis in the concept of saturation level. The solubility product used to predict the pH at saturation (pH s ) for a water is empirically modified in the Stiff-Davis index. The Stiff-Davis index will predict that a water is less scale forming than the LSI calculated for the same water chemistry and conditions. The deviation between the indices increases with ionic strength. Interpretation of the index is by the same scale as for the Langelier saturation index. Oddo-Tomson index. The Oddo-Tomson index accounts for the impact of pressure and partial pressure of CO 2 on the pH of water and on the sol- ubility of calcium carbonate. This empirical model also incorporates cor- rections for the presence of two or three phases (water, gas, and oil). Interpretation of the index is by the same scale as for the LSI and Stiff- Davis indices. Momentary excess (precipitation to equilibrium). The momentary excess index describes the quantity of scalant that would have to precipitate instantaneously to bring water to equilibrium. In the case of calcium carbonate, K spc ϭ [Ca 2ϩ ] [CO 3 2Ϫ ] If water is supersaturated, then [Ca 2ϩ ] [CO 3 2Ϫ ] ӷK spc Precipitation to equilibrium assumes that one mole of calcium ions will precipitate for every mole of carbonate ions that precipitates. On this basis, the quantity of precipitate required to restore water to equi- librium can be estimated with the following equation: [Ca 2ϩ Ϫ X] [CO 3 2Ϫ Ϫ X] ϭ K spc where X is the quantity of precipitate required to reach equilibrium. X will be a small value when either calcium is high and carbonate low, or carbonate is high and calcium low. It will increase to a maximum when equal parts of calcium and carbonate are present. As a result, these calculations will provide vastly different values for waters with the same saturation level. Although the original momentary excess index was applied only to calcium carbonate scale, the index can be extended to other scale-forming species. In the case of sulfate, momen- tary excess is calculated by solving for X in the relationship 110 Chapter Two 0765162_Ch02_Roberge 9/1/99 4:02 Page 110 [Ca 2ϩ Ϫ X] [SO 4 2Ϫ Ϫ X] ϭ K spc The solution becomes more complex for tricalcium phosphate: [Ca 2ϩ Ϫ 3X] 3 [PO 4 3Ϫ Ϫ 2X] 2 ϭ K spc While this index provides a quantitative indicator of scale poten- tial and has been used to correlate scale formation in a kinetic mod- el, the index does not account for two critical factors: First, the pH can often change as precipitates form, and second, the index does not account for changes in driving force as the reactant levels decrease because of precipitation. The index is simply an indicator of the capacity of water to scale, and can be compared to the buffer capaci- ty of a water. Interpreting the indices. Most of the indices discussed previously describe the tendency of a water to form or dissolve a particular scale. These indices are derived from the concept of saturation. For example, saturation level for any of the scalants discussed is described as the ratio of a compound’s observed ion-activity product to the ion-activity product expected if the water were at equilibrium K sp . The following general guidelines can be applied to interpreting the degree of super- saturation: 1. If the saturation level is less than 1.0, a water is undersaturated with respect to the scalant under study. The water will tend to dis- solve, rather than form, scale of the type for which the index was calculated. As the saturation level decreases and approaches 0.0, the probability of forming this scale in a finite period of time also approaches 0. 2. A water in contact with a solid form of the scale will tend to dissolve or precipitate the compound until an IAP/K sp ratio of 1.0 is achieved. This will occur if the water is left undisturbed for an infi- nite period of time under the same conditions. A water with a satu- ration level of 1.0 is at equilibrium with the solid phase. It will not tend to dissolve or precipitate the scale. 3. As the saturation level (IAP/K sp ) increases above 1.0, the tenden- cy to precipitate the compound increases. Most waters can carry a moderate level of supersaturation before precipitation occurs, and most cooling systems can carry a small degree of supersatu- ration. The degree of supersaturation acceptable for a system varies with parameters such as residence time, the order of the scale reaction, and the amount of solid phase (scale) present in the system. Environments 111 0765162_Ch02_Roberge 9/1/99 4:02 Page 111 2.2.4 Ion association model The saturation indices discussed previously can be calculated based upon total analytical values for all possible reactants. Ions in water, however, do not tend to exist totally as free ions. 24 Calcium, for example, may be paired with sulfate, bicarbonate, carbonate, phosphate, and oth- er species. Bound ions are not readily available for scale formation. This binding, or reduced availability of the reactants, decreases the effective ion-activity product for a saturation-level calculation. Early indices such as the LSI are based upon total analytical values rather than free species primarily because of the intense calculation requirements for determining the distribution of species in a water. Speciation of a water requires numerous computer iterations for the following: 25 ■ The verification of electroneutrality via a cation-anion balance, and balancing with an appropriate ion (e.g., sodium or potassium for cation-deficient waters; sulfate, chloride, or nitrate for anion-defi- cient waters). ■ Estimating ionic strength; calculating and correcting activity coef- ficients and dissociation constants for temperature; correcting alkalinity for noncarbonate alkalinity. ■ Iteratively calculating the distribution of species in the water from dissociation constants. A partial listing of these ion pairs is given in Table 2.13. ■ Verification of mass balance and adjustment of ion concentrations to agree with analytical values. ■ Repeating the process until corrections are insignificant. ■ Calculating saturation levels based upon the free concentrations of ions estimated using the ion association model (ion pairing). The ion association model has been used by major water treatment companies since the early 1970s. The use of ion pairing to estimate the concentrations of free species overcomes several of the major shortcom- ings of traditional indices. While indices such as the LSI can correct activity coefficients for ionic strength based upon the total dissolved solids, they typically do not account for common ion effects. Common ion effects increase the apparent solubility of a compound by reducing the concentration of available reactants. A common example is sulfate reducing the available calcium in a water and increasing the apparent solubility of calcium carbonate. The use of indices which do not account for ion pairing can be misleading when comparing waters in which the TDS is composed of ions which pair with the reactants and of ions which have less interaction with them. 112 Chapter Two 0765162_Ch02_Roberge 9/1/99 4:02 Page 112 The ion association model provides a rigorous calculation of the free ion concentrations based upon the solution of the simultaneous non- linear equations generated by the relevant equilibria. 26 A simplified method for estimating the effect of ion interaction and ion pairing is sometimes used instead of the more rigorous and direct solution of the equilibria. 27 Pitzer coefficients estimate the impact of ion association upon free ion concentrations using an empirical force fit of laboratory data. 28 This method has the advantage of providing a much less calcu- lation-intensive direct solution. It has the disadvantages of being based upon typical water compositions and ion ratios, and of unpre- dictability when extrapolated beyond the range of the original data. The use of Pitzer coefficients is not recommended when a full ion asso- ciation model is available. When indices are used to establish operating limits such as maxi- mum concentration ratio or maximum pH, the differences between indices calculated using ion pairing can have some serious economic significance. For example, experience on a system with high-TDS water may be translated to a system operating with a lower-TDS water. The high indices that were found acceptable in the high-TDS water may be Environments 113 TABLE 2.13 Examples of Ion Pairs Used to Estimate Free Ion Concentrations Aluminum [Aluminum] ϭ [Al 3ϩ ] ϩ [Al(OH) 2ϩ ] ϩ [Al(OH) 2 ϩ ] ϩ [Al(OH) 4 Ϫ ] ϩ [AlF 2ϩ ] ϩ [AlF 2 ϩ ] ϩ [AlF 3 ] ϩ [AlF 4 Ϫ ] ϩ [AlSO 4 ϩ ] ϩ [Al(SO 4 ) 2 Ϫ ] Barium [Barium] ϭ [Ba 2ϩ ] ϩ [BaSO 4 ] ϩ [BaHCO 3 ϩ ] ϩ [BaCO 3 ] ϩ [Ba(OH) ϩ ] Calcium [Calcium] ϭ [Ca 2ϩ ] ϩ [CaSO 4 ] ϩ [CaHCO 3 ϩ ] ϩ [CaCO 3 ] ϩ [Ca(OH) ϩ ] ϩ [CaHPO 4 ] ϩ [CaPO 4 Ϫ ] ϩ [CaH 2 PO 4 ϩ ] Iron [Iron] ϭ [Fe 2ϩ ] ϩ [Fe 3ϩ ] ϩ [Fe(OH) ϩ ] ϩ [Fe(OH) 2ϩ ] ϩ [Fe(OH) 3 Ϫ ] ϩ [FeHPO 4 ϩ ] ϩ [FeHPO 4 ] ϩ [FeCl 2ϩ ] ϩ [FeCl 2 ϩ ] ϩ [FeCl 3 ] ϩ [FeSO 4 ] ϩ [FeSO 4 ϩ ] ϩ [FeH 2 PO 4 ϩ ] ϩ [Fe(OH) 2 ϩ ] ϩ [Fe(OH) 3 ] ϩ [Fe(OH) 4 Ϫ ] ϩ [Fe(OH) 2 ] ϩ [FeH 2 PO 4 2ϩ ] Magnesium [Magnesium] ϭ [Mg 2ϩ ] ϩ [MgSO 4 ] ϩ [MgHCO 3 ϩ ] ϩ [MgCO 3 ] ϩ [Mg(OH) ϩ ] ϩ [MgHPO 4 ] ϩ [MgPO 4 Ϫ ] ϩ [MgH 2 PO 4 ϩ ] ϩ [MgF ϩ ] Potassium [Potassium] ϭ [K ϩ ] ϩ [KSO 4 Ϫ ] ϩ [KHPO 4 Ϫ ] ϩ [KCl] Sodium [Sodium] ϭ [Na ϩ ] ϩ [NaSO 4 Ϫ ] ϩ [Na 2 SO 4 ] ϩ [NaHCO 3 ] ϩ [NaCO 3 Ϫ ] ϩ [Na 2 CO 3 ] ϩ [NaCl] ϩ [NaHPO 4 Ϫ ] Strontium [Strontium] ϭ [Sr 2ϩ ] 1 [SrSO 4 ] 1 [SrHCO 3 ϩ ] ϩ [SrCO 3 ] ϩ [Sr(OH) ϩ ] 0765162_Ch02_Roberge 9/1/99 4:02 Page 113 unrealistic when translated to a water where ion pairing is less signif- icant in reducing the apparent driving force for scale formation. Table 2.14 summarizes the impact of TDS upon LSI when it is calculated using total analytical values for calcium and alkalinity, and when it is calculated using the free calcium and carbonate concentrations deter- mined with an ion association model. Indices based upon ion association models provide a common denom- inator for comparing results between systems. For example, calcite saturation level calculated using free calcium and carbonate concen- trations has been used successfully as the basis for developing models which describe the minimum effective scale inhibitor dosage that will maintain clean heat-transfer surfaces. 29 The following cases illustrate some practical usage of the ion association model. Optimizing storage conditions for low-level nuclear waste. Storage costs for low-level nuclear wastes are based upon volume. Storage is therefore most cost-effective when the aqueous-based wastes are concentrated to occupy the minimum volume. Precipitation is not desirable because it can turn a low-level waste into a high-level waste, which is much more costly to store. Precipitation can also foul heat-transfer equipment used in the concentration process. The ion association model approach has been used at the Oak Ridge National Laboratory to predict the optimum conditions for long-term storage. 30 Optimum conditions involve the parameters of maximum concentration, pH, and temperature. Figures 2.17 and 2.18, respectively, depict a profile of the degree of supersatura- tion for silica and for magnesium hydroxide as a function of pH and tem- perature. It can be seen that amorphous silica deposition may present a problem when the pH falls below approximately 10, and that magne- sium hydroxide or brucite deposition is predicted when the pH rises above approximately 11. Based upon this preliminary run, a pH range of 10 to 11 was recommended for storage and concentration. Other potential precipitants can be screened using the ion association model to provide an overall evaluation of a wastewater prior to concentration. 114 Chapter Two TABLE 2.14 Impact of Ion Pairing on the Langelier Scaling Index (LSI) LSI Water Low TDS High TDS TDS impact on LSI High chloride No pairing 2.25 1.89 Ϫ0.36 With pairing 1.98 1.58 Ϫ0.40 High sulfate No pairing 2.24 1.81 Ϫ0.43 With pairing 1.93 1.07 Ϫ0.86 0765162_Ch02_Roberge 9/1/99 4:02 Page 114 Limiting halite deposition in a wet high-temperature gas well. There are several fields in the Netherlands that produce hydrocarbon gas asso- ciated with very high TDS connate waters. Classical oilfield scale problems (e.g., calcium carbonate, barium sulfate, and calcium sul- fate) are minimal in these fields. Halite (NaCl), however, can be pre- cipitated to such an extent that production is lost in hours. As a result, a bottom-hole fluid sample is retrieved from all new wells. Unstable components are “fixed” immediately after sampling, and pH is determined under pressure. A full ionic and physical analysis is also carried out in the laboratory. The analyses were run through an ion association model computer program to determine the susceptibility of the brine to halite (and other scale) precipitation. If a halite precipitation problem was predicted, the ion association model was run in a “mixing” mode to determine if mixing the connate water with boiler feedwater would prevent the problem. This Environments 115 25 34 43 53 62 71 80 13 12 11 10 9 8 7 0 2 4 6 8 10 12 14 16 Saturation Level Temperature pH Figure 2.17 Amorphous silica saturation in low-level nuclear wastewater as a function of pH and temperature (WaterCycle). 0765162_Ch02_Roberge 9/1/99 4:02 Page 115 approach has been used successfully to control salt deposition in the well with the composition outlined in Table 2.15. The ion association model evaluation of the bottom-hole chemistry indicated that the water was slightly supersaturated with sodium chloride under the bottom-hole con- ditions of pressure and temperature. As the fluids cooled in the well bore, the production of copious amounts of halite was predicted. The ion association model predicted that the connate water would require a minimum dilution with boiler feedwater of 15 percent to pre- vent halite precipitation (Fig. 2.19). The model also predicted that over- injection of dilution water would promote barite (barium sulfate) formation (Fig. 2.20). Although the well produced H 2 S at a concentra- tion of 50 mg/L, the program did not predict the formation of iron sul- fide because of the combination of low pH and high temperature. Boiler feedwater was injected into the bottom of the well using the downhole 116 Chapter Two 25 34 43 53 62 71 80 13 12 11 10 9 8 7 0 5000 10000 15000 20000 25000 30000 35000 40000 45000 Saturation Level Temperature pH Figure 2.18 Brucite saturation in low-level nuclear wastewater as a function of pH and temperature. 0765162_Ch02_Roberge 9/1/99 4:02 Page 116 injection valve normally used for corrosion inhibitor injection. Injection of dilution water at a rate of 25 to 30 percent has allowed the well to produce successfully since start-up. Barite and iron sulfide precipita- tion have not been observed, and plugging with salt has not occurred. Identifying acceptable operating range for ozonated cooling systems. It has been well established that ozone is an efficient microbiological con- trol agent in open recirculating cooling-water systems (cooling towers). It has also been reported that commonly encountered scales have not been observed in ozonated cooling systems under conditions where scale would otherwise be expected. The water chemistry of 13 ozonat- ed cooling systems was evaluated using an ion association model. Each system was treated solely with ozone on a continuous basis at the rate of 0.05 to 0.2 mg/L based upon recirculating water flow rates. 31 Environments 117 25 42 58 75 92 108 125 100 83 67 50 33 17 0 0 0.5 1 1.5 2 2.5 Degree of Saturation Temperature % Injection Figure 2.19 Degree of saturation of halite in a hot gas well as a function of temperature and reinjected boiler water (DownHole SAT). 0765162_Ch02_Roberge 9/1/99 4:02 Page 117 The saturation levels for common cooling-water scales were calcu- lated, including calcium carbonate, calcium sulfate, amorphous silica, and magnesium hydroxide. Brucite saturation levels were included because of the potential for magnesium silicate formation as a result of the adsorption of silica upon precipitating magnesium hydroxide. Each system was evaluated by 31 ■ Estimating the concentration ratio of the systems by comparing recirculating water chemistry to makeup water chemistry. ■ Calculating the theoretical concentration of recirculating water chemistry based upon makeup water analysis and the apparent, cal- culated concentration ratio from step 1. ■ Comparing the theoretical and observed ion concentrations to deter- mine precipitation of major species. ■ Calculating the saturation level for major species based upon both the theoretical and the observed recirculating water chemistry. ■ Comparing differences between the theoretical and actual chem- istry to the observed cleanliness of the cooling systems and heat exchangers with respect to heat transfer surface scale buildup, scale formation in valves and on non–heat-transfer surfaces, and precipitate buildup in the tower fill and basin. 118 Chapter Two TABLE 2.15 Hot Gas Well Water Analysis Bottom hole connate Boiler feedwater Temperature, °C 121 70 Pressure, bars 350 1 pH, site 4.26 9.10 Density, kg/m 3 1.300 1.000 TDS, mgиL Ϫ1 369,960 Ͻ20 Dissolved CO 2 , mgиL Ϫ1 223 Ͻ1 H 2 S (gas phase), mgиL Ϫ1 50 0 H 2 S (aqueous phase), mgиL Ϫ1 Ͻ0.5 0 Bicarbonate, mgиL Ϫ1 16 5.0 Chloride, mgиL Ϫ1 228,485 0 Sulfate, mgиL Ϫ1 320 0 Phosphate, mgиL Ϫ1 Ͻ10 Borate, mgиL Ϫ1 175 0 Organic acids ϽC 6 , mgиL Ϫ1 12 Ͻ5 Sodium, mgиL Ϫ1 104,780 Ͻ1 Potassium, mgиL Ϫ1 1,600 Ͻ1 Calcium, mgиL Ϫ1 30,853 Ͻ1 Magnesium, mgиL Ϫ1 2,910 Ͻ1 Barium, mgиL Ϫ1 120 Ͻ1 Strontium, mgиL Ϫ1 1,164 Ͻ1 Total iron, mgиL Ϫ1 38.0 Ͻ0.01 Lead, mgиL Ϫ1 5.1 Ͻ0.01 Zinc, mgиL Ϫ1 3.6 Ͻ0.01 0765162_Ch02_Roberge 9/1/99 4:02 Page 118 [...]... 28 88 483 216 238 4 95 549 480 496 317 2 308 2623 610 A† 36 38 16 8 223 320 607 13 5 78 78 50 8 303 2972 20 Silica ⌬‡ T* A† ⌬‡ System cleanliness Ϫ8 50 3 15 Ϫ7 Ϫ82 11 2 414 402 418 2664 5 Ϫ349 59 0 40 24 38 66 11 2 16 2 11 2 280 18 6 3 050 12 6 610 0 19 52 52 20 31 48 10 1 14 3 10 1 78 60 95 12 6 13 8 85 12 4 7 18 11 19 11 202 12 6 29 95 0 59 62 18 67 No scale observed Basin buildup Heavy scale Valve scale Condenser tube... Page 12 0 1 (1) 2 (2) 3 (2) 4 (2) 5 (3) 6 (3) 7 (3) 8 (3) 9 (3) 10 (3) 11 (3) 12 (3) 13 (3) Theoretical vs Actual Recirculating Water Chemistry 9 /1/ 99 4:02 TABLE 2 .16 076 51 6 2_Ch02_Roberge System (Category) Calcite Brucite Silica T* A† T* A† T* A† 0.03 49 89 10 6 240 54 0 59 8 794 809 11 98 16 70 3420 7634 0.02 5. 4 611 50 72 51 28 26 6 .5 62 74 37 65 Ͻ0.0 01 0.82 2.4 1. 3 3.0 5. 3 10 53 10 7.4 4.6 254 7.6 Ͻ0.0 01. .. Ni Al bronze 70/30 Cu/Ni ϩ Fe Type 316 stainless steel 6% Mo stainless steel Ni-Cu alloy 40 Quiet seawater 2.0 4.9 0. 25 0.32 Nil 1. 12 0. 25 1. 8 Nil 1. 3 0.0 75 0 .55 0.027 0. 017 0.02 0. 055 Ͻ0.02 0.02 0. 01 0.02 8.2 mиs 1 35 42 mиs 1 — 4.4 0.9 1. 8 0.2 0.22 0 .12 Ͻ0.02 Ͻ0.02 Ͻ0. 01 4 .5 13 .2 1. 07 1. 32 0.97 0.97 1. 47 Ͻ0. 01 Ͻ0. 01 0. 01 076 51 6 2_Ch02_Roberge 14 2 9 /1/ 99 4:02 Page 14 2 Chapter Two that this declines... The corresponding formula is given in Eq (2 . 15 ).36 S ϭ 0.0080 Ϫ 0 .16 92K0 .5 ϩ 25. 3 853 K ϩ 14 .0941K1 .5 Ϫ 7.0261K2 ϩ 2.7081K2 .5 (2 . 15 ) 076 51 6 2_Ch02_Roberge 9 /1/ 99 4:02 Page 13 1 Environments 13 1 Note that in this definition, (‰) is no longer used, but an old value of 35 corresponds to a new value of 35 Since the introduction of this practical definition, salinity of seawater is usually determined by measuring... function of temperature and reinjected boiler water 076 51 6 2_Ch02_Roberge 12 0 System (Category) Calcium T* A† 56 80 238 288 392 803 14 64 800 7 75 3904 417 0 3660 7930 43 60 288 18 0 2 45 16 3 200 16 8 95 270 18 8 800 68 *T ϭ theoretical (ppm) †A ϭ actual (ppm) ‡⌬ ϭ difference (ppm) Magnesium ⌬‡ 13 20 50 10 8 14 7 640 12 64 632 680 3634 3982 2860 7862 T* 28 88 483 216 238 4 95 549 480 496 317 2 308 2623 610 A† 36 38 16 8... 0.02 5. 4 611 50 72 51 28 26 6 .5 62 74 37 65 Ͻ0.0 01 0.82 2.4 1. 3 3.0 5. 3 10 53 10 7.4 4.6 254 7.6 Ͻ0.0 01 0.02 0 .12 0 .55 0.46 0.73 0 .17 0.06 Ͻ0. 01 0.36 0.36 0 .59 0 .14 0.20 0.06 0 .10 0 .13 0. 21 0. 35 0.40 0 .10 0.22 0. 31 0.22 1. 31 1.74 0. 25 0.09 0 .12 0 .16 0. 35 0.49 0 .52 0.33 0.27 0. 35 0.44 0 .55 0 .10 *T ϭ theoretical (ppm) †A ϭ actual (ppm) Observation No scale observed Basin buildup Heavy scale Valve scale... temperature T (K) and salinity S (‰) are known: 35 ln [O2] (mL и L 1) ϭ A1 ϩ A2 (10 0/T) ϩ A3 ln (T /10 0) ϩ A4 (T /10 0) ϩ S[B1 ϩ B2 (T /10 0) ϩ B3 (T /10 0)2] 076 51 6 2_Ch02_Roberge 13 4 9 /1/ 99 4:02 Page 13 4 Chapter Two where A1 ϭ 17 3.4292 A2 ϭ 249.6339 A3 ϭ 14 3.3483 A4 ϭ Ϫ 21. 8492 B1 ϭ Ϫ0.033096 B2 ϭ 0. 014 259 B3 ϭ Ϫ0.0 017 000 The primary source of the dissolution of oxygen is the air-sea exchange with oxygen in... mmol 1 kg 1 gиkg 1 Naϩ 468 .5 10 . 21 53 .08 10 .28 0.090 54 5.9 0.842 0.068 2.30 28.23 0. 416 10 .77 0.399 1. 290 0. 412 1 0.0079 19 . 354 0.0673 0.0 013 0 .14 0 2. 712 0.0 257 Kϩ Mg2ϩ Ca2 ϩ Sr2 ϩ ClϪ BrϪ FϪ HCO3Ϫ SO42Ϫ B(OH)3 is converted to oxide, and when all organic matter is completely oxidized The definition of 19 02 was translated into Eq (2 .13 ), where the salinity (S) and chlorinity (Cl) are expressed in parts per... Silica 076 51 6 2_Ch02_Roberge 12 6 9 /1/ 99 4:02 Page 12 6 Chapter Two 350 Soluble silica (mg/l) 300 250 200 15 0 10 0 50 8.9 8.6 8.4 0 56 50 7.6 44 38 Temperature 32 pH 7 26 20 Figure 2. 21 Solubility of amorphous silica as a function of temperature and pH TABLE 2 .19 Silica Limits for Three Treatment Schemes Low pH (6.0) Temperature (°C) Silica level (ppm) Saturation level limit 20 13 0 1. 2 30 15 0 1. 1 Moderate... limit for this pH range has been 15 0 ppm as SiO2 As outlined in Table 2 .19 , a limit of 15 0 ppm would correspond roughly to a saturation level of 1. 4 at 20°C and 1. 1 at 30°C At the upper end of the cooling-water pH range (9.0), silica solubility increases to 11 5 ppm (20°C) and 14 0 ppm (30°C) A control limit of 18 0 ppm would correspond to saturation levels of 1. 5 and 1. 3, respectively In systems where . 18 Valve scale 5 (3) 392 2 45 14 7 238 320 Ϫ82 11 2 10 1 11 Condenser tube scale 6 (3) 803 16 3 640 4 95 607 11 2 16 2 14 3 19 No scale observed 7 (3) 14 64 200 12 64 54 9 13 5 414 11 2 10 1 11 No scale observed 8. mgиL 1 10 Borate, mgиL 1 1 75 0 Organic acids ϽC 6 , mgиL 1 12 5 Sodium, mgиL 1 104,780 1 Potassium, mgиL 1 1,600 1 Calcium, mgиL 1 30, 853 1 Magnesium, mgиL 1 2, 910 1 Barium, mgиL 1 120. water flow rates. 31 Environments 11 7 25 42 58 75 92 10 8 12 5 10 0 83 67 50 33 17 0 0 0 .5 1 1 .5 2 2 .5 Degree of Saturation Temperature % Injection Figure 2 .19 Degree of saturation of halite in a hot

Ngày đăng: 05/08/2014, 09:20

Từ khóa liên quan

Tài liệu cùng người dùng

  • Đang cập nhật ...

Tài liệu liên quan