APRIL 1998 ZHANG ET AL. 589

Sensitivity of the GFDL Modular Model to Parameterization of Double-Diffusive Processes*

JUBAO ZHANG MIT/WHOI Joint Program, Department of , Woods Hole Oceanographic Institution, Woods Hole, Massachusetts

RAYMOND W. S CHMITT AND RUI XIN HUANG Department of Physical Oceanography, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts (Manuscript received 24 February 1997, in ®nal form 7 July 1997)

ABSTRACT The effect of double-diffusive mixing on the general circulation is explored using the GFDL MOM2 model. The motivation for this comes from the known sensitivity of the thermohaline circulation to the vertical diffusivity and the earlier work of Gargett and Holloway, who studied the effects of a simple nonunity ratio between heat and salt diffusivities in a GCM. In this work, a more realistic, yet conservative, parameterization of the double-

diffusive mixing is applied, with the intensity depending on the local density ratio R␳ ϭ ␣Tz/␤Sz. A background diffusivity is used to represent non-double-diffusive turbulent mixing in the stably strati®ed environment. The numerical model is forced by relaxation boundary conditions on both temperature and salinity at the sea surface. Three control experiments have been carried out: one with the double-diffusive parameterization (DDP) deter- mined by the local density ratio, one with constant but different diffusivities for heat and salt as previously considered by Gargett and Holloway (GHD), and the other with the same constant diapycnal eddy diffusivity for both heat and salt (CDD). The meridional overturning in run DDP is 22% less than in run CDD, and the maximum poleward heat transport is about 8% less. In comparison, the overturning rate and poleward heat transport in run GHD display reductions that are about half as large. The interior temperature and salinity in run DDP and GHD are higher than in run CDD, with the change in run DDP more than twice that in run GHD. In addition, in DDP and GHD, the density ratio distribution becomes closer to unity than in run CDD, with the change in run DDP being larger than in GHD. Interestingly, the double diffusion is stronger in the western boundary current region than the interior, implying a close relation between vertical shear and the intensity of double diffusion. These results indicate a greater sensitivity of the thermohaline circulation to double diffusion than had previously been suspected due to the tendency of the double-diffusive mixing to generate self-reinforcing ¯ows. This effect appears to be more signi®cant when the double-diffusive mixing is applied only when the strati®cation is favorable rather than uniformly applied. In addition, parameter sensitivity experiments suggest that double diffusion could have stronger effects on the meridional overturning and poleward heat transport than modeled here since the parameterizations chosen are rather conservative.

1. Introduction pycnal) eddy diffusivity (K), which is the mechanism that provides the necessary warming of the rising abys- The thermohaline circulation of the ocean plays an sal waters. F. Bryan (1987) found an approximate K 1/3 important role in the climate system of the earth, es- dependence of the MOC; however, his model runs may pecially in controlling the oceanic part of the poleward not have reached equilibrium; later work has found a heat transport. The large-scale structure of the ther- K 2/3 dependence (Marotzke 1997; Zhang et al. 1997, mohaline circulation and its variability has been studied submitted to J. Phys. Oceanogr.), when the conventional extensively in models (Weaver and Hughes 1992; Huang ``relaxation'' boundary conditions are applied for the 1995). One result of these studies is the ®nding that the surface forcing. Whatever the speci®c power law de- strength of the meridional overturning cell (MOC) is pendence, the continual production of cold deep water strongly dependent on the value of the vertical (or dia- at high latitudes requires vertical mixing to close the circulation (Munk 1966). Observational microstructure work in recent years * WHOI Contribution Number 9239. (Gregg 1989; Polzin et al. 1995) has indicated that tur- bulent mixing in the ocean interior caused by internal Corresponding author address: Mr. Jubao Zhang, Room 54-1419, wave breaking is generally much weaker than has been Massachusetts Institute of Technology, Cambridge, MA 02139. inferred from the large-scale budget approach of Munk E-mail: [email protected] (1966), Hogg et al. (1982), and others. However, evi-

᭧ 1998 American Meteorological Society

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 590 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28 dence is emerging that turbulence near rough topogra- plied in the climate-scale numerical model runs de- phy may be suf®ciently enhanced to provide the re- scribed below. However, the parameterization of mixing quired downward heat ¯ux (Toole et al. 1994; Polzin et processes, both vertical and horizontal, requires careful al. 1997). One problem with near-bottom mixing is that treatment in numerical models. The oceanic circulation the strati®cation is weak in the abyss, so heat ¯uxes involves extremely broad scales in both time and space; tend to be small, even with very large eddy diffusivities. thus, we can never resolve all the temporal and spatial This reduced effectiveness of deep internal-wave-in- scales in numerical models. Consequently, for the fore- duced turbulence (the mixing of already mixed water) seeable future, subgrid-scale phenomena must be par- leads us to examine the other mixing processes occur- ameterized in ocean general circulation models. In fact, ring in the more strongly strati®ed thermocline. determining suitable parameterizations of subgrid-scale The primary processes that must be considered are phenomena is one of the most critical problems in nu- the double-diffusive instabilities of salt ®ngering, which merical modeling of the general circulation. Early ocean occurs when temperature and salinity both decrease with circulation models were mostly based on z coordinates, depth, and diffusive convection, which occurs when and subgrid-scale mixing was parameterized in terms of temperature and salinity both increase with depth. It is constant horizontal and vertical mixing. However, such well known that on the molecular level heat and salt simple horizontal/vertical mixing schemes can introduce have diffusivities that are two orders of magnitude dif- strong arti®cial cross-isopycnal mixing near fronts, such ferent. This difference drives convective motions even as the Gulf Stream, where isopycnals slope sharply. To if the overall density pro®le is stable. Substantial evi- overcome the arti®cial cross-isopycnal mixing in low- resolution z-coordinate models, a numerical technique dence is now available that these processes play a sig- of rotating the mixing tensor has been proposed by Redi ni®cant role in ocean mixing (Schmitt 1994). Their prin- (1982) and has been implemented here. cipal effect is to transport heat and salt at different rates In addition, Gent and McWilliams (1990, hereafter in the vertical and cause a net upgradient ¯ux of buoy- GM90) pointed out that there is an extra advection term ancy. A number of recent modeling studies show that in the tracer balance equation of non-eddy-resolving such differential transports have important effects on models, whose existence is due to the Lagrangian mean the stability and structure of the large-scale thermoha- transport of mesoscale eddies. They introduced a pa- line circulation. rameterization based on a downgradient diffusion of the In particular, Gargett and Holloway (1992, hereafter isopycnal layer thickness in adiabatic ¯ow. The addition GH92) ®rst investigated the effects of a simple double- of this transport term makes the tracer conservation diffusive parameterization in an ocean model. They equations in a non-eddy-resolving model self-consistent showed that the steady-state characteristics of low-res- and eliminates the unphysical background horizontal olution GCMs used in climate studies are very sensitive diffusion require in the earlier version of the isopycnal/ to the ratio of the vertical eddy diffusivities for salinity diapycnal mixing schemes. The dynamic effect and and temperature. Large differences in meridional trans- model sensitivity on the isopycnal/diapycnal mixing ports resulted due to the upgradient buoyancy ¯ux, have been discussed in several recent papers (Gent et which forced different advective±diffusive balances to al. 1995; Danabasoglu and McWilliams 1995). be realized. Ruddick and Zhang (1989) found that the In this study we will focus on the parameterization ``salt oscillator'' mechanism in a box model could be of diapycnal mixing in connection with double-diffusive completely stabilized by incorporating salt ®ngers into processes. Based on results from theory, laboratory ex- the model. Most recently, Gargett and Ferron (1996) periments, and oceanic observations, the intensity of found that multibox models with nonequal heat and salt double diffusion must depend on the local density ratio diffusivities exhibited extended ranges of multiple equi- R␳ ϭ ␣Tz/␤Sz. There is strong evidence indicating that libria, a different mode transition near present-day val- double diffusion is important in controlling the diapyc- ues of freshwater forcing magnitude, and the possibility nal mixing process in the ocean only if R␳ is suf®ciently of quasiperiodic oscillatory states compared to an equal close to one. Thus, we propose here a conservative pa- diffusivity run. Climate models generally neglect the rameterization of double diffusion that re¯ects this de- double-diffusive processes, so there is a need for ex- pendence (section 2). This parameterization is applied panded studies of their effects. to the GFDL MOM2 (Pacanowski 1995) model. The Since the Gargett and Holloway and Gargett and Fer- numerical experiments and results are presented and in- ron parameterizations of double diffusion are rather sim- terpreted in section 3. In addition, sensitivity study ex- ple, one must be concerned that their studies might over- periments are presented in section 4 in order to under- estimate the impact of differential mixing. There is good stand the response of the numerical model to the double- evidence that the double-diffusive processes are most diffusive parameterization, and a summary follows in active only under certain hydrographic conditions. Thus, section 5. as a next step in exploring the role of double diffusion, we review the scienti®c literature related to double-dif- 2. Parameterization of double-diffusive processes fusive processes and propose a more realistic parame- Double-diffusive convection can occur in a stably terization of double diffusion for OGCMs. This is ap- strati®ed environment where the vertical density gra-

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 591 dient due to either temperature or salinity is destabiliz- tic to the Gulf Stream in the Florida Straits (Schmitz et ing. For the case with warm, salty water on top of cold al. 1993). freshwater, salt ®ngering occurs; for the case with cold In addition to the strong vertical mixing inferred for freshwater on top of warm, salty water, diffusive lay- the staircase regions such as the tropical Atlantic, Med- ering appears (Turner 1973; Stern 1975; Schmitt 1994). iterranean out¯ow, and the Tyrhennian Sea, where the Double diffusion occurs because of the two orders of density ratio is less than 1.7, ®ngers are also thought to magnitude difference in the molecular diffusivities of play some role in mixing in the broad central waters of heat and salt. Double-diffusive mixing is driven by the the World Ocean, where the density ratio is only slightly release of potential energy in the redistribution of the higher. Some contribution is very likely since turbulence destabilizing component, which has the larger buoyancy due to shear instability is infrequent, occurring only a ¯ux. A signi®cant fraction of the released energy is used few percent of the time (Polzin 1996). Laboratory ex- to move the stable component upward. However, the periments have shown that ®ngers grow rapidly after a total density ¯ux is downward, so the overall potential turbulent mixing event, when the density ratio is near energy of the water column is reduced. As a result, the common oceanic value of 2 (Taylor 1991). When double-diffusive processes are self-energized through ®ngers alone exist, there is a strong increase in labo- the change in the mean potential energy. In the case of ratory ¯uxes as the density ratio drops below 2.0 (Mc- ordinary turbulent mixing, however, the potential energy Dougall and Taylor 1984). Gargett and Schmitt (1982) of the water column is increased; thus external energy ®nd microscale signatures of ®ngers in the Central Wa- sources such as internal waves or strong mean shears ters of the North Paci®c. Mack (1985, 1989) and Mack are required. and Schoeberlein (1993) ®nd salt ®nger signatures in Conditions favorable for double diffusion are very microstucture data from the upper thermocline of the common in the . In the subtropics the ocean gains Sargasso Sea when the local density ratio falls below heat from the atmosphere, and evaporation exceeds pre- 3.0. Marmorino et al. (1987) also ®nds the distinct spec- cipitation. Thus, in comparison with the deep ocean, the tral signatures of salt ®ngers in data from the seasonal upper ocean is warm and salty, a strati®cation which thermocline of the Sargasso Sea. Recent theoretical favors salt ®ngering. Ingham (1966) reports that 90% work of Shen and Schmitt (1995) provides a salt ®nger of the main thermocline of the Atlantic Ocean has a model spectrum that is in good agreement with the ob- density ratio less than 2.3. Schmitt (1990) ®nds that 95% served microstructure. It is based on the ``dispersion of the upper kilometer in the Atlantic at 24ЊN is salt relation'' for salt ®ngers derived by Schmitt (1979), ®ngering favorable. In the western tropical North At- which shows that ®nger growth times are comparable 2 lantic an area of over 1 million km was found to contain to the local buoyancy period only when R␳ is less than strong thermohaline staircases in the main thermocline 2. Hamilton et al. (1989) report on salt ®nger signatures (Schmitt et al. 1987). Systematic, large-scale water mass in microstructure data from the main thermocline of the changes within the staircase layers were observed, eastern North Atlantic that appear to provide a signif- which cannot be explained by conventional mechanical icant vertical nutrient ¯ux. turbulence but ®t nicely the enhanced salt ¯ux expected The extensive observational, laboratory, and theoret- from salt ®ngers. This location is characterized by a ical evidence that ®ngers are most intense when the strong vertical salt gradient and a low value of the den- density ratio is less than about 2 leads us to propose a sity ratio R␳. Microstructure observations (Gregg and parameterization that is dependent on R␳. Schmitt Sanford 1987; Lueck 1987; Marmorino et al. 1987) re- (1981) has argued that the constancy of the large-scale vealed narrowband, limited amplitude structures of the density ratio observed in the main thermocline can be predicted scale for salt ®ngers within the interfaces, and explained by this dependence and the greater transfer plumelike convective structures within the mixed layers. rate for salt. In the modeling below we adopt a param- The intensity of the microstructure indicates that an eterization similar to his [and to the form inferred by eddy diffusivity for salt of 1±2 cm2 sϪ1 must apply with- Kunze (1990)], though reduced in amplitude to conform in the staircase (Schmitt 1988; Kunze 1990). While low- with recent observations. er than expected from the simple application of labo- Due to surface cooling and excess precipitation, the ratory ¯ux laws, this is quite substantial compared to surface water in polar and subpolar regions is colder the weak background value of 0.07 cm2 sϪ1 predicted and fresher than the subsurface water. As a result, the for the canonical Garret and Munk (1972) internal wave other type of double diffusion, diffusive convection (or ®eld (Gregg 1989). Indeed, the vertical heat ¯ux derived layering), is possible. In this case, the heat content is from the microstructure measurements is nearly 3 W the destabilizing component and more heat is mixed mϪ2, which exceeds the ¯ux estimated in a very strongly upward than salt. Though the initial process may appear turbulent, but weakly strati®ed, abyssal site near the as an oscillatory overstability (Stern 1960; Shirtcliffe Mid-Atlantic Ridge (Polzin et al. 1997). The salt ®nger 1967), the fully developed state is one of alternating mixing in this region of the Atlantic is thought to be convecting layers and high gradient sheets (Turner responsible for the water mass conversion observed in 1973); hence the . The characteristic staircase lay- the thermocline waters transiting from the South Atlan- ers have long been observed beneath the ice in the Arctic

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 592 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28

Ocean (Neal et al. 1969; Neshyba et al. 1971). Padman where KS and KT are the diapycnal eddy diffusivities and Dillon (1987, 1988) provide evidence for wide- for salinity and temperature, respectively; K ϱ is the con- spread diffusive staircases in both the Beaufort Sea and stant diapycnal diffusivity due to other mixing processes Canadian Basin. The Antarctic contains sites of strong unrelated to double diffusion, such as internal wave diffusive layering as well. Muench et al. (1990) found breaking, which is independent of R␳; and Rc is the numerous thermohaline staircases in the northwestern critical density ratio above which the diapycnal mixing Weddell Sea during March 1986. Throughout an ap- due to salt ®ngering drops dramatically, due to the ab- proximately 40 000 km2 study area, they observed that sence of staircases. A value of 0.7 is used for the heat/ between depths of 100 and 180 m, staircases with a salt buoyancy ¯ux ratio due to salt ®ngers. Here K*is typical scale of 1±5 m occurred almost everywhere. Be- the maximum diapycnal diffusivity due to salt ®ngers. tween depths of 180 and 500 m, over ϳ50% of the We have chosen a more modest value than originally study area were staircases with layer thickness of 10± proposed by Schmitt (1981), re¯ecting improved un- 100 m. Vertical heat ¯uxes of up to 15 W mϪ2 were derstanding of ¯uxes in thermohaline staircases ob- estimated, which is signi®cant in the upper-ocean heat served in the C-SALT program (Schmitt 1988). Expo- budget and may help maintain ice-free conditions. Rob- nent n is an index to control the decay of KT, KS with ertson et al. (1995) also found that diapycnal ¯uxes in increasing R␳. the Weddell Sea were dominated by diffusive convec- In numerical models, vertical gradients of tempera- tion. Similarly, Martinson (1990) found it to be impor- ture and salinity, Tz and Sz, are not well simulated com- tant in a model of winter sea ice formation. Kelley pared to the oceanic observations. The simulation is (1984, 1990) has provided a parameterization of the especially poor in the deep ocean where the vertical mixing due to diffusive convection, which derives from gradients of temperature and salinity in low-resolution a successful scaling of the layer thicknesses in both the numerical models are very small and often noisy. On laboratory and ocean. Effective diffusivities are weaker the other hand, most ®eld observations of double dif- than in the salt ®nger regime, consistent with the con- fusion are limited to the upper ocean, and it is not clear vective transports realized with ®ngers and the conduc- whether double diffusion can play any signi®cant role tive ¯uxes obtained in diffusive interfaces. at great depth where the vertical gradients of both tem- In traditional numerical models diapycnal mixing in perature and salinity are small. Thus, we have intro- the stably strati®ed part of the model ocean is para- duced an additional constraint that double diffusion can meterized in terms of the same constant for both heat occur only if the magnitude of the vertical temperature and salt. In this study we will divide the stably strati®ed gradient is larger than a critical value: region into three subregions: |Tz| Ͼ Tz,c. (3) 1) salt ®ngering regime 2) diffusive layering regime This is a conservative step to limit salt ®ngering to the 3) non-double-diffusive regime. thermocline, where evidence indicates its importance, and avoid it in the abyss, where such evidence is not The convective adjustment scheme for regions of un- yet available. stable strati®cation remains the same as in the normal MOM2 code. b. Diffusive layering regime a. Salt ®ngering regime When Salt ®ngers can form when warmer saltier water over- T Ͻ 0, S Ͻ 0, 1 Ͼ R Ͼ k /k , (4) lies colder fresher water such that z z ␳ s t T Ͼ 0, S Ͼ 0, 1 Ͻ R Ͻ k /k , (1) diffusive layering can occur. z z ␳ t s Kelley (1984) discussed a parameterization for dif- where R␳ ϭ ␣Tz/␤Sz is the density ratio, and kt and ks fusive layering process in which he applied the labo- are the molecular diffusivities of heat and salt respec- ratory-derived double-diffusive ¯ux laws to the oceanic tively. The ratio of molecular diffusivities is about 100, data. The formulation is given as follows: so a very weak destabilizing salinity gradient can induce K CR1/3 k salt ®ngering. However, observations indicate that ®n- Tatϭ (5) gers become signi®cant mixing agents in the ocean only K ϭ RRK, when the density ratio is close to unity (Schmitt 1994). SF␳ T Here we apply the parameterization of Schmitt (1981) where

K* Ϫ0.54(RϪ1Ϫ1) K ϭϩK ϱ C ϭ 0.00859 exp{4.6e ␳ }, S n 1 ϩ (R␳/Rc) 9 Ϫ1.1 Ra ϭ 0.25 ϫ 10 R␳ , (6) 0.7K* ϱ K ϭϩK , (2) and kt is the molecular diffusivity of heat; RF is the T n R␳␳[1 ϩ (R /Rc)] buoyancy ¯ux ratio de®ned as

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 593

␤F explicitly. Instead, K is de®ned diagnostically as a func- R ϭ S , (7) ␳ F ␣F tion of the density ratio, assuming a locally linear equa- T tion of state applies to the mixing between adjacent in which FS, FT are upward ¯ux of salt and heat re- layers. An expression equivalent to that given in GH92 spectively. Huppert's (1971) formulation for RF has been is readily obtained: used in Kelley (1984): RK␳ TSϪ K K␳ ϭ . (12) 1.85 Ϫ 0.85/R␳␳, 0.5 Յ R Ͻ 1.0 R Ϫ 1 R ϭ (8) ␳ F 0.15, R Ͻ 0.5. Ά ␳ The change of diffusivities of temperature, salinity, 2 Ϫ1 Roughly speaking, KS ഠ KT ഠ 1.0 cm s when R␳ and density with the density ratio is shown in Fig. 1, → 4 1.0. Here KS and KT decrease approximately as R␳ where we take the double-diffusive parameters as 2 andR␳ , respectively, over the range 0.1 Յ R␳ Յ 1.0. 2 Ϫ1 ϱ 2 Ϫ1 Furthermore, Kelley (1990) re®ned his formulation K* ϭ 1.0 cm s , K ϭ 0.3 cm s , by giving a new empirical formulation of C(R␳) and RF: Rc ϭ 1.6, n ϭ 6, 0.72 C ϭ 0.0032 exp{4.8R } Ϫ4 Ϫ1 ␳ Tz,c ϭ 2.5 ϫ 10 ЊC m . (13) 3/2 1/R␳␳ϩ 1.4(1/R Ϫ 1) Because heat and salt mixing rates are different, the R ϭ . (9) F 3/2 1 ϩ 14(1/R␳ Ϫ 1) equivalent buoyancy diffusivity varies greatly from one region to another (Fig. 1): Similar to the case of salt ®ngering, KT and KS are different and they depend on the density ratio. However, 1) In the weak double-diffusive regime (R␳ Ͻ 0.3 or the diapycnal diffusivity suggested by Kelley (1984, R␳ Ͼ 3.0), K␳ is close to its upper limit, given by

1990) was rather weak at low density ratios since KS the constant turbulent mixing coef®cient. Ϫ3 fell below the molecular diffusivity for heat (1.4 ϫ 10 2) In the moderate double diffusion regime (0.3 Ͻ R␳ 2 Ϫ1 cm s ) when R␳ Ͻ 0.25, which could cause compu- Ͻ 0.77 or 0.97 Ͻ R␳ Ͻ 1.0 or 1.56 Ͻ R␳ Ͻ 3.0), tational dif®culty in a numerical model. On the other K␳ is reduced, so the mixing of buoyancy is less hand, given the role of non-double-diffusive processes ef®cient. in the diapycnal mixing (Muench et al. 1990), a back- 3) In the strong double-diffusion regime (0.77 Ͻ R␳ Ͻ ground diffusivity is also needed in the region of dif- 0.97 or 1.0 Ͻ R␳ Ͻ 1.56), the sign of K␳ is reversed. fusive layering. Thus, we include K ϱ in the parameter- Thus, upgradient buoyancy diffusion appears be- ization for diffusive layering: cause the buoyancy ¯ux due to the destabilizing component is larger than that of the stabilizing com- K CR1/3 k K ϱ Tatϭ ϩ ponent. ϱϱ KSFϭ RR␳(KT Ϫ K ) ϩ K , (10) 3. Control experiments where C and RF are given in (9) and Ra de®ned in (6). Also, the restraint (3) is applied in the diffusive layering a. The numerical experiments case as well. Note, however, the de®nition of density ratio in this paper is inverse to that in Kelley (1984, The GFDL MOM2 is used in these experiments. The 1990) in order to deal with both forms of double dif- details of the model can be found in K. Bryan (1969) fusion. His formulation is quite conservative compared and Pacanowski (1995). The speci®c features of these to that of Fedorov (1988) or Muench (1990), with weak- experiments are as follows: er ¯uxes when R␳ is close to 1. The model domain is a rectangular basin of 60Њ by 60Њ, and the horizontal resolution is 3.75Њ. There are 15 levels vertically, with thickness of 50 m, 75 m, 100 m, c. Non-double-diffusive regime 125 m, 150 m, 200 m, 250 m, 300 m, 400 m, 450 m, Away from the regions discussed above, we use a 450 m, 450 m, 500 m, 500 m, and 500 m from the top constant background diapycnal mixing rate for both to the bottom, and the total depth is 4500 m. tracers to represent internal-wave-induced mixing: The annual zonally averaged climatological surface wind stress (Hellerman and Rosenstein 1993), SST, and K ϭ K ϭ K ϱ. (11) T S SSS (Levitus 1982) are used in these experiments. The The sinking limb of the thermohaline circulation is ®rst layer temperature and salinity are relaxed to the driven by buoyancy losses at the air±sea interface; the SST and SSS climatology, with a relaxation time of 30 return limb involves a balance between upward buoy- days. ancy advection and downward buoyancy diffusion in The horizontal and vertical momentum viscosity co- 10 2 Ϫ1 2 the ocean interior. However, in the GFDL MOM2, den- ef®cients are Ah ϭ 1.0 ϫ 10 cm s and A␷ ϭ 20 cm sity is a diagnostic variable; thus the diapycnal eddy sϪ1 respectively. For the tracer equations, the eddy trans- diffusivity of density K␳ does not appear in the model port parameterization of GM90 and isopycnal/diapycnal

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 594 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28

FIG. 1. The change of diapycnal eddy diffusivities of temperature (solid line), salinity (broken

line), and density (dotted line) with density ratio R␳. In the left half panel, where R␳ Ͻ 1, diffusive

layering mixing occurs. In comparison, in the right half panel where R␳ Ͼ 1, salt ®ngers form. Note

that even though diffusive layering region and salt ®ngering region seem to be continuous at R␳ ϭ 1, in the real ocean, normally there is a gravitationally unstable transition region between them

where R␳ Ͻ 0. mixing is used, and no background horizontal diffusion that everything is identical except for the parameter- 6 2 Ϫ1 is needed. We take KISO ϭ KITD ϭ 5 ϫ 10 cm s , ization of diapycnal diffusion. All experiments run from where KISO is the isopycnal diffusion coef®cient and KITD the same initial conditions and under the same boundary is the downgradient diffusivity of the isopycnal thick- conditions for 8000 years. For the reason of compara- ness. bility, we calculate the basin-averaged KT and KS in run 2 Ϫ1 2 Ϫ1 Three control experiments will be discussed here, the DDP: K T ϭ 0.326 cm s and K S ϭ 0.366 cm s , 2 Ϫ1 ®rst with the double-diffusive parameterization (DDP and then we use KT ϭ 0.326 cm s and KS ϭ 0.366 2 Ϫ1 2 Ϫ1 hereafter) for temperature and salinity, and the second cm s in run GHD and KT ϭ KS ϭ 0.346 cm s in with the diffusivity given in a similar fashion to GH92 run CDD; therefore the average diffusivities of both (Gargett and Holloway diffusivity, GHD hereafter), that scalars are equal in all three experiments. is, the diffusivities for heat and salt are different but A technical issue is the computational cost of adding both kept constant everywhere, and the third with the on the double-diffusive parameterization. For a 100-yr conventional assumption of KS ϭ KT ϭ const, (constant integration on an SGI workstation, the CPU time is diapycnal diffusivity, CDD hereafter). Gargett and Hol- 1256, 1120, and 1113 CPU seconds for runs DDP,GHD, loway (1992) discussed many experiments in their pa- and CDD respectively. Thus, the double-diffusive pa- per, but given the differences in the formulation of the rameterization takes approximately 10% extra compu- numerical model, particularly the inclusion of GM90, tational time, a rather modest cost. we deemed it worthwhile to run an experiment similar to theirs and to compare the results directly. The parameters in the double-diffusion parameteriza- b. Results and interpretations tion are de®ned in (13). Note that this parameterization is more conservative than that of Schmitt (1981). This 1) DIFFUSIVITY VARIABILITY IN RUN DDP is adopted because results of the C-SALT observations (Schmitt 1988) indicate that a modest reduction of the The major difference between DDP and GHD/CDD mixing coef®cient was appropriate. The uncertainties is that KT and KS in run DDP depend on the spatial associated with these parameters and vertical resolution variability of density ratio R␳. It is found that salt ®ngers will be discussed in detail in the next section. are the dominant double-diffusive phenomenon; thus, We designed these three experiments in such a way here we only map out the distribution of the salt dif-

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 595

FIG.2.TheKS distribution in run DDP of different sections: (a) horizontal section in the upper thermocline (350 m); (b) horizontal section in the lower thermocline (950 m); (c) zonal section in the subtropical gyre (32ЊN); (d) zonal section in the subpolar gyre (50ЊN); (e) meridional section close to the western boundary (305ЊE); and (f) meridional section in the interior (334ЊE). The contour interval is 0.05 cm2 sϪ1.

fusivity and the corresponding KT distribution can be The enhanced salt ®ngering is associated with strong deduced accordingly. vertical salinity gradients within the western bound- To visualize the complicated three-dimensional struc- ary region. According to the classic picture of the ture of the circulation we plot several two-dimensional oceanic general circulation there are surface and maps of KS in Fig. 2. deep western boundary currents. The surface western There are two distinct features in the KS distribution: boundary current brings warm and salty water north- 1) The double diffusion is con®ned within the upper ward, and the deep western boundary current brings 1500 m; thus only the upper 2000 m are plotted in cold and fresh water southward along the western Figs. 2c±f. Double diffusion is stronger near the bot- boundary. As a result, the temperature and salinity tom of the thermocline (e.g., 950 m in Fig. 2b) than distributions within the western boundary region are in shallower layers (e.g., 350 m in Fig. 2a). The lack dominated by horizontal advection, which maintains of double-diffusive activity in the deep part of the strong vertical gradients of salinity along the path of model ocean is due to the temperature gradient con- the western boundary currents, and this is the region straint (3), which eliminates the double-diffusive most favorable for salt ®ngering because R␳ is processes in the deep ocean where the vertical gra- brought closer to 1. In comparison, vertical diffusion dients in the numerical model are weak. This con- may play a more prominent role in the oceanic in- straint also re¯ects that we have no reliable obser- terior, thus limiting the vertical gradient of salinity vational data to support double-diffusive phenomena there. As a result, salt ®ngering in the ocean interior in the deep ocean. Such a constraint can be modi®ed is not as strong as within the western boundary re- in the future based on new observations. gion where vertical shear between poleward and

2) In the western boundary region KS is stronger than equatorward water masses maintains the salinity in the interior, indicating strong salt ®ngering there. contrast.

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 596 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28

FIG. 4. Relation between K␳ and R␳ in the three different experi- ments.

formation at high latitudes. The deep water returns to the source region through basinwide uniform upwelling in the ocean interior. To reach an equilibrium, the dia- pycnal advection (upwelling) must be compensated by the diapycnal diffusion. Thus, the magnitude of the ther- mohaline circulation is very sensitive to the diapycnal diffusivity of density, as shown by F. Bryan (1987). Given the importance of the diapycnal diffusivity of

density, we plot the relation of K␳ and R␳ in Fig. 4. In run CDD, K␳ ϭ KS ϭ KT is constant; however, in run GHD and DDP, K␳ depends strongly on R␳. In the region FIG. 3. The zonally integrated meridional overturning volume of diffusive layering (R␳ Ͻ 1), K␳ in run GHD is larger streamfunctions for run (a) DDP, (b) GHD, and (c) CDD. Contour than the constant K and K speci®ed in the experiment. interval is 0.5 Sv. T S This is contrary to the common belief that diffusive layering reduces or even reverses the vertical buoyancy diffusion, as in run DDP. In GH92, this drawback was 2) MERIDIONAL OVERTURNING avoided by giving d ϭ KS/KT Ͼ 1ord ϭ KS/KT Ͻ 1 The thermohaline circulation can be visualized by according to whether salt ®ngering or diffusive layering plotting the zonally integrated meridional overturning is favorable; thus, in both regimes K␳ is either reduced streamfunction. The structure of the thermohaline cir- or with the sign reversed, in accordance with double- culation in the three cases (Fig. 3) is similar, but the diffusive processes. magnitudes are signi®cantly different. The overturning In the experiments studied here, salt ®ngering is the rate in run CDD is 6.17 Sv (1 Sv ϵ 106 m3 sϪ1); while dominant double-diffusive process; therefore, only a it is only 4.79 Sv in run DDP, a 22% decrease compared very small region in run GHD is affected by an un- with run CDD. In run GHD the meridional overturning physical representation of the diffusive layering process. is 5.55 Sv, a 10% decrease compared with run CDD. The circulation is primarily controlled by salt ®ngering, There are traces of two-grid noise near the equator in and the relation between the density ratio and density all three experiments, which may be due to the coarse diffusivity is plotted in the right half of Fig. 4. We can vertical resolution (Weaver and Sarachik 1990). In ad- see that both GHD and DDP are consistent with the dition, a very weak reversed meridional cell in the equa- ®ngering mechanism while displaying signi®cant quan- torial deep ocean appears in run GHD and CDD. How- titative differences. First of all, given the same R␳, K␳ ever, there is no reversed cell in DDP, and this may be is largest in run CDD, and then GHD, and smallest in due to the fact that the double-diffusion parameteriza- run DDP. This difference increases when R␳ is closer to tion helps to eliminate the local water column instability, 1. As we will discuss later, the implementation of a which leads to the local reversed cell. double-diffusive parameterization has an effect on the According to the classic Stommel±Arons theory distribution of density ratio (Turner angle), which leads (Stommel and Arons 1960), deep water moves equa- to a wider area of double-diffusively favorable condi- torward as a deep western boundary current after its tions. Thus, we can imagine that the difference between

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 597

meridional overturning rate multiplied by the temper- ature difference between the surface and deep water. While the maximum of the thermohaline circulation lies at high latitudes, the largest surfaceϪdeep water tem- perature difference appears at lower latitudes. Thus, the maximum poleward heat transport is found in midlati- tudes, where the percentage of reduction of the ther- mohaline circulation at that latitude is less than that of the maximum overturning rate, and the temperature dif- ference remains nearly the same for two experiments. Second, the wind-driven gyre and Ekman cell also con- tribute to the poleward heat transport; these are con- trolled by the surface wind stress and remain almost the same for both experiments.

4) TURNER ANGLE FIG. 5. The poleward heat transport for the three experiments.

Since R␳ can vary from Ϫϱ to ϱ and it cannot rep- resent the convective overturning regime, we would like the model runs may be greater than the diffusivity dif- to describe the spatial structure of the density ratio by ferences at the same R␳. mapping the Turner angle (Tu). Ruddick (1983; 1996 Since downward buoyancy diffusion in run DDP is personal communication) de®ned the Turner angle as weaker than in runs GHD and CDD, a smaller amount the four quadrant arctangent of R␳, that is, of upwelling is needed to balance the weaker density diffusion. Thus, the magnitude of the thermohaline cir- R ϩ 1 Tu ϭ arctan␳ . (14) culation in run DDP is reduced. Furthermore, upwelling R Ϫ 1 is not spatially uniform. In fact, upwelling within the ΂΃␳ western boundary region is two orders of magnitude Thus, salt ®ngering occurs for 45ЊϽTu Ͻ 90Њ, and stronger than in the interior in all the runs (Huang and it is strongest when Tu is close to 90Њ. In comparison, Yang 1996); thus, upwelling within the western bound- diffusive layering occurs for Ϫ90ЊϽTu ϽϪ45Њ and ary region constitutes an essential part of the upward when Tu is close to Ϫ90Њ, diffusive layering is strong. limb of the thermohaline circulation. As discussed in When Ϫ45ЊϽTu Ͻ 45Њ, the ¯uid is in the non-double- section 2, when double diffusion is suf®ciently strong diffusive regime. Other Tu values are within the grav-

(R␳ close to 1), buoyancy diffusion is reduced or re- itationally unstable region. versed. Strong double diffusion in the western boundary The Turner angle maps from the three experiments region (Fig. 2) dramatically reduces the downward are complex; Tu at level 7 (950 m) and level 13 (3500 buoyancy diffusion and can reverse its direction locally. m) is plotted in Fig. 6. Differences among the three Even though diapycnal diffusion in the ocean interior experiments are obvious: (i) Near the bottom of the is less affected, the total amount of the diapycnal dif- thermocline (level 7), salt ®ngering favorable conditions fusive ¯ux of buoyancy is reduced and, therefore, the are found in all experiments. Interestingly, the Tu in run thermohaline circulation becomes weaker. DDP is highest, and then run GHD, and lowest in run CDD, indicating that salt ®nger mixing in DDP is re- inforced, especially near the western boundary current 3) POLEWARD HEAT TRANSPORT (WBC). This also occurs in run GHD, though not as With signi®cant changes in the magnitude of the ther- strongly as in run DDP. (ii) In the deep ocean (level 13), mohaline circulation, it is not surprising that the pole- the Turner angle is only marginally critical in limited ward heat transport is also affected by the different pa- regions in run CDD, but salt ®ngering (in the subtropical rameterizations of diapycnal processes (Fig. 5). The gyre) and diffusive layering (in subpolar gyre) both oc- poleward heat transport at all latitudes is weakest in run cur in run DDP. In run GHD, however, only the salt DDP and strongest in run CDD, while it is intermediate ®ngering favorable condition is found, and it is weaker in run GHD. The difference in poleward heat ¯ux is than in run DDP. most signi®cant at midlatitudes. The maximum pole- Thus, double diffusion appears to change the distri- ward heat ¯ux in run CDD is 0.255 PW, and it is 0.246 bution of the Turner angle. In these model runs, this PW in run GHD and 0.236 PW in run DDP, a 4% and parameterization renders the water column even more 8% decrease respectively. Note that the decrease in the favorable to double diffusion. Although both salt ®n- heat transport is less than the decrease in the thermo- gering and diffusive layering are possible in run DDP, haline circulation. This can be explained as follows. strong salt ®ngering appears over most of the basin, First, the poleward heat transport is proportional to the while diffusive layering is con®ned to narrow regions

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 598 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28

FIG. 6. The Turner angle at the bottom of level 7 (950 m) for run (a) DDP, (b) GHD, (c) CDD and at the bottom of level 13 (3500 m) for run (d) DDP (e) GHD, and (f) CDD respectively. Note that uneven contour intervals have been used in the above ®gure.

in the subpolar gyre near the surface or close to the hancing because of the weakening of the destabilizing bottom. The simple representation of double diffusion gradient, which would shift R␳ away from one (Schmitt in GHD makes the deep water more salt ®ngering fa- 1981). However, in a two- or three-dimensional model, vorable but eliminates the regime for diffusive layering double-diffusive activity can reinforce itself. The dou- as in run DDP and CDD. ble-diffusive lateral intrusions described by Turner The density ratio is the most important index for (1978), Ruddick and Turner (1979), and Ruddick double-diffusive activity. It controls the intensity of (1992) are a clear example of such self-reinforcing double diffusion and thus affects the large-scale ther- ¯ows. That is, the maintenance and intensi®cation of mohaline circulation and water mass formation (see the double-diffusively favorable density ratio are due next). For both double-diffusive runs, we found that to the action of vertical shear on isopycnal gradients. the distribution of density ratio in the upper kilometer The shear arises from horizontal pressure gradients de- at 24ЊN was in better agreement with the hydrographic veloped when the double-diffusive ¯uxes begin af- data for the Atlantic as reported by Schmitt (1990) than the constant diffusivity run. However, the improvement fecting the vertical density pro®le. Whereas thermo- was small since the relaxation boundary conditions haline intrusions have vertical scales of 100 m or less, play a dominant role in setting upper-ocean T±S struc- it appears that such effects occur with much larger ture. We ®nd that the addition of double diffusion caus- vertical scale within our model as well, based on the es the density ratio to become closer to one, which is comparison of the DDP and CDD runs. As there are somewhat counterintuitive (Fig. 6). This means that 1000-m vertical-scale water masses in the ocean with double-diffusive processes and large-scale circulation similarly low density ratios, it would be interesting to interact in a complicated way. Double diffusion in a determine how much of it is due to self-enhancing one-dimensional (vertical) situation cannot be self-en- effects versus the juxtaposition of shear and isopycnal

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 599

FIG. 7. Vertical pro®les of horizontally averaged (a) temperature and (b) salinity, respectively, where the solid line is for run DDP, the dotted broken line for run GHD, and the broken line for run CDD. The difference from run CDD of (c) temperature and (d) salinity respectively. gradients due to the wind- and buoyancy-driven cir- main thermocline are dictated by vertical advection and culations. vertical diffusion with the horizontal advection playing a minor role only. Under the relaxation conditions for both temperature 5) TEMPERATURE AND SALINITY DISTRIBUTION and salinity, surface temperature and salinity are nearly The double-diffusive parameterization is also found ®xed to their climatological mean values. Since the sa- to affect the basinwide water mass properties. Changes linity diffusivity in run DDP and GHD is larger than in in the temperature and salinity distribution due to DDP run CDD, salinity is higher due to stronger downward can be seen from the horizontally averaged temperature salt diffusion in both DDP and GHD, compared with and salinity pro®les (Fig. 7). Compared to run CDD, CDD, as shown in Fig. 7. the basin-averaged temperature in run DDP and GHD Changes in the temperature pro®le can be explained is lower in the uppermost 200 m, while it is higher below by a simple one-dimensional advection±diffusion bal- 200 m. On the other hand, the horizontally averaged ance, neglecting the lateral advection and diffusion salinity in both DDP and GHD is higher for the whole terms. That is, the temperature distribution is controlled water column. The difference from run CDD is similar by the following one-dimensional balance wTz ϭ ␬Tzz, for run GHD and DDP, and both display the most sig- where z is the distance along the path connecting the ni®cant differences in the main thermocline. On the oth- cold source and the hot source. Assuming both w and er hand, the difference from CDD seems twice as large ␬ are constant along the path, the solution of this one- in run DDP as in run GHD for the whole depth. The dimensional equation is in form of T ϭ T1 ϩ C(T 2 Ϫ az aD basin averaged temperature is 6.069ЊC for run DDP, T1)(e Ϫ 1), where C ϭ 1/(e Ϫ 1) in which D is the 5.973ЊC for run GHD, and 5.937ЊC for run CDD, and distance between the cold and hot source and a ϭ w/␬ the salinity is 33.749 psu, 33.723 psu, and 33.711 psu, is the inverse of the characteristic scale of this advec- respectively; thus, water in run DDP is 0.13ЊC warmer tion±diffusion problem. Small a means a diffusion-dom- and 0.04 psu saltier than in run CDD, while it is 0.04ЊC inated case, with almost a linear pro®le between the warmer and 0.01 psu saltier in run GHD compared to cold and heat source, while large a indicates a sharp run CDD. front, like the main thermocline. Tracer distributions are controlled by advection and Salt mixing (circulation) is stronger (weaker) in run diffusion. While the wind-driven horizontal advection DDP and GHD, compared with CDD. Thus, the salinity plays a prominent role in setting up the tracer distri- pro®le in these two cases should have a broad halocline, bution in the upper ocean, tracer distributions below the thus a high salinity in the abyssal water. From this so-

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 600 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28

TABLE 1. Parameter sensitivity experiments and results.

2 Ϫ1 2 Ϫ1 2 Ϫ1 K* (cm s ) Rc n THC (Sv) KT (cm s ) KS (cm s ) 1.0 1.6 6 4.793 0.326 0.366 0.0 1.6 6 5.578 0.300 0.300 0.2 1.6 6 5.505 0.304 0.310 0.5 1.6 6 5.137 0.311 0.329 2.0 1.6 6 4.156 0.350 0.426 5.0 1.6 6 3.475 0.379 0.506 1.0 1.2 6 5.245 0.305 0.313 1.0 2.0 6 3.935 0.356 0.441 1.0 1.6 2 3.759 0.353 0.436 1.0 1.6 12 5.252 0.310 0.323 1.0 1.6 32 5.310 0.306 0.313 2.0 1.2 6 5.176 0.327 0.351 0.5 2.0 6 4.650 0.329 0.375 lution, it is clear that strong diffusion and weak circu- processes and the other part, K ϱ, represents the con- lation give rise to a broad main thermocline and warm ventional, non-double-diffusive mixing due to turbu- and salty abyssal water, as shown in Fig. 7. lence (from internal wave breaking). The magnitudes of The temperature pro®le is slightly more complicated both K* and K ϱ are not ®rmly established, and it is to explain. In run DDP, strong mixing above the bottom worthwhile to examine the sensitivity of the model to of the main thermocline transports more heat downward, the ratio K*/K ϱ. For simplicity, we ®x K ϱ ϭ 0.3 cm2 so temperature is higher in the main thermocline, as sϪ1 and change the value of K*. shown in Fig. 7. Warm water in the main thermocline This background diffusivity is higher than the value makes the water in the whole basin warmer than in run inferred from the tracer release experiment of Ledwell CDD, even though thermal diffusion below the main et al. (1993), which suggested a diffusivity of 0.15 cm2 thermocline in run DDP is weaker than in run CDD. sϪ1 in the main thermocline (which may be partially The shallow cold temperature anomaly near the surface explained by salt ®ngers). As Yin and Fung (1991) have in run DDP re¯ects that thermal diffusion there is lower shown, unphysical vertical mixing in models with un- than average. even vertical grids may occur, if the diffusivity is low, The temperature pro®le in run GHD is the result of and this contaminates the numerical results. In order to a delicate balance between a weakened circulation and avoid such numerical problems in our low-resolution weakened diffusion. Compared with run CDD, the cir- experiments, we have used a slightly high value of 0.3 culation in run GHD is 10% weaker and diffusion is cm2 sϪ1, which represent a compromise among the af- 5% weaker. Thus, the characteristic scale 1/a is reduced, fordable vertical resolution, isopycnal mixing rate, and and diffusion becomes more important than advection. the observations. Therefore, we only need to test three As a result, the main thermocline is slightly broader, parameters, and the experiments and results are sum- thus the warm anomaly compared with run CDD, as marized in Table 1. shown in Fig. 7. From Table 1 we can see that different parameters can lead to very different results. When we increase K*, 4. Sensitivity experiments that is, the contribution from the double-diffusive part, both K and K increase in a quasi-linear fashion and Although the parameterization of double diffusion T S used here is consistent with existing theoretical, labo- the increase of double-diffusive processes decreases the ratory, and observational results, the parameterization diapycnal diffusion of density; thus the magnitude of involves several free parameters. Therefore, it is im- the meridional overturning is reduced. Similarly when portant to test the sensitivity of the model runs to these we increase Rc or decrease n, allowing modest ®ngering parameters as well as other speci®cs of the numerical to affect a greater portion of the water column, the mean model, such as the vertical resolution. Note that all pa- diffusivity for heat and for salt increases and the ther- rameters in these experiments are the same as in run mohaline circulation weakens. DDP unless stated otherwise. The relationship between R␳ and the eddy diffusivities for salt and heat is still not ®xed. Schmitt (1981) applied the laboratory-derived ¯ux laws to the observational a. Parameter sensitivity data and estimated the eddy diffusivities for salt and There are four variables in the parameterization given heat. The maximum diffusivity due to salt ®ngers could ϱ 2 Ϫ1 in section 2: K*, K , Rc, and n. One special feature of reach 5.0 cm s or more, however, given the uncer- this parameterization is that diapycnal mixing consists tainty associated with the empirical ¯ux laws and more of two parts: one part accounts for the double-diffusive understanding gained through the C-SALT (Schmitt et

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 601

TABLE 2. Tz,c sensitivity experiments and results.

Ϫ4 Ϫ1 2 Ϫ1 2 Ϫ1 Tz,c (10 ЊCm ) THC (Sv) KT (cm s ) KS (cm s ) 0.0 4.733 0.386 0.476 1.0 4.749 0.337 0.389 2.5 4.811 0.326 0.366 5.0 4.861 0.323 0.360 10.0 4.944 0.318 0.348 al. 1987) ®eld program, we judged that a good ®rst step in the distribution of temperature and salinity. In com- was to use a somewhat conservative value for K*, like paring a DDP run with Tz,c ϭ 0.0 to a CDD experiment 2 Ϫ1 2 Ϫ1 K* ϭ 1.0 cm s . As even this conservative value leads with KT ϭ KS ϭ 0.43 cm s (the mean diffusivity in to a marked difference in the thermohaline circulation, the DDP run), we ®nd that the temperature is lower in it suggests that more work should to be done to under- the main thermocline (0±500 m) and warmer below 500- stand the role of salt ®ngers in oceanic mixing. m depth, and the salinity is higher for the whole water

Both Rc and n affect the relation among KT, KS, and depth. These changes in water mass properties caused

R␳, and little is known about the exact value of them. by the double diffusion are of interest for climate stud- However, on the basis of the limited observational data, ies. We do not intend to compare results from our ide-

KS appears to be large for R␳ less than 1.5, falling dra- alized runs with observations. However, we hope that matically near R␳ ϭ 1.7 and reaching background levels the underlying dynamics should apply to the more prac- by R␳ ϭ 1.9. Thus, Rc ϭ 1.6, n ϭ 6 are reasonable tical and complicated climate models. Danabasoglu and based on observation. Even though the theoretical con- McWilliams (1995, DM hereafter) have tested the sen- dition for salt ®ngers is 1 Ͻ R␳ Ͻ 100, observational sitivity of a global ocean circulation model to param- data suggests that signi®cant ®ngering only occurs when eterizations of mesoscale isopycnal tracer transports and

R␳ is close enough to 1. A similar conclusion holds for showed substantial improvement in several climatically the diffusive layering. Therefore, a relatively stronger important aspects of the ocean circulation, compared vertical gradient of salt (temperature) is required for the with the conventional horizontal/vertical eddy diffusion salt ®ngering (diffusive layering) to be important in the parameterization. However, the horizontally averaged diapycnal mixing processes. temperature is warmer in the main thermocline and colder in the deep ocean compared with Levitus data (Levitus 1982) in their model (refer to Fig. 8 in DM). b. Sensitivity to restraint T z,c The salinity is fresher throughout the water column,

Experiments and results with different Tz,c are sum- especially in the deep ocean (refer to Fig. 16 in DM). marized in Table 2. The magnitude of Tz,c can make a These discrepancies are just the opposite of the changes signi®cant difference in basin-averaged diffusivities for we see in our Tz,c ϭ 0 run. Thus, our results suggest salt and heat, while the magnitude of meridional over- that the vertical temperature and salinity distributions turning remains nearly the same. The reason is that the could be improved in such global climate models if a effect of the constraint Tz,c works like a ®lter: it only parameterization of double diffusion were to be applied affects the part of ocean below the main thermocline in the deep ocean. In this regard we note that McDougall where the vertical gradients of temperature and salinity and Whitehead (1984) invoked some salt ®ngering to are weak. In the main thermocline, however, Tz is much explain the evolution of water mass properties in the larger than the Tz,c given in the above table, and thus Antarctic Bottom Water in the Atlantic. the double diffusion in the main thermocline remains almost the same. As found by Cummins et al. (1990), c. Salt ®ngering and diffusive layering alone an increase of diffusivity in the deep ocean (they use a diffusivity that is inversely dependent on the local buoy- To study the different roles played by salt ®ngers and ancy frequency) increases the vertical gradient of T, S diffusive layering in run DDP, we ran two experiments in the deep ocean, but the magnitude of the thermohaline in which these processes are separated with only one of circulation is barely changed. them active in each experiment. For example, in the salt

The criteria of a minimum Tz,c for double diffusion ®ngering alone run, the salt ®ngering parameterization to occur is speculative. We viewed it as a way to avoid (2) is activated, but diffusive layering is shut off; that ϱ 2 Ϫ1 spuriously large double-diffusion in the weakly strati- is, a constant KT ϭ KS ϭ K ϭ 0.3 cm s is applied ®ed abyss, where there is little evidence for its impor- in places with no salt ®nger activity. The experiment tance, and noise in the computation of vertical gradients for diffusive layering only is designed similarly. might be a problem. In the salt ®ngering alone experiment, the meridional Even though the double-diffusive activity in the deep overturning rate is 4.80 Sv, and the globally averaged 2 Ϫ1 ocean does not affect the magnitude of thermohaline diffusivities are K T ϭ 0.325 cm s and K S ϭ 0.366 circulation much, it can make a signi®cant difference cm2 sϪ1, which is very close to run DDP. In contrast,

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 602 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28 in the diffusive layering alone experiment the meridi- this new experiment is 5.02 Sv, and the basin-averaged 2 Ϫ1 2 onal overturning rate is 5.58 Sv, and both K T and K S diffusivities are K T ϭ 0.337 cm s , K S ϭ 0.388 cm is no more than 0.0003 cm2 sϪ1 higher than K ϱ. There- sϪ1. Compared to run DDP, it is clear that both the fore, diffusive-layering activity is barely noticeable in thermohaline circulation and basin-average mixing rates these runs. Even though changes in KT and KS are almost K T, K S intensi®ed. trivial, the meridional overturning rate is enhanced com- The increase of the strength of the thermohaline cir- pared to the overturning rate of 5.575 Sv for the case culation can be attributed to two aspects: (i) The increase with K* ϭ 0, as shown in Table 1. This result implies of vertical resolution alone can increase the magnitude that diffusive layering may tend to intensify the ther- of the thermohaline circulation. We ran two experiments 2 Ϫ1 mohaline circulation. This might occur because the pro- with constant KT ϭ KS ϭ 0.45 cm s , one with 15 cess can release heat to the upper ocean without ex- vertical levels and another with 30 levels. The meridi- periencing deep convection, thereby converting a warm, onal overturning rate was 7.47 Sv and 7.64 Sv respec- salty water mass to a cold, salty water mass that more tively, with a small difference of 0.17 Sv. Thus, we can readily sinks. McDougall (1983) invokes such a mech- expect a similar effect in the DDP experiments. (ii) Even anism for Greenland Sea Bottom Water formation. How- though the averaged KT, KS are larger for the case with ever, for our model, the effect is small, possibly because the higher vertical resolution, the strength of double of the 60Њ northern limit to the domain and the lack of diffusion in the main thermocline is actually slightly an Arctic Ocean, where the most extensive diffusive lowered. Since the increase of vertical resolution can layering is found (Padman and Dillon 1988). increase the vertical gradient of temperature and salinity Both experiments show that under restoring boundary (Weaver and Sarachik 1990), especially in the deeper conditions, salt ®nger mixing occurs much more widely parts of the ocean, double diffusion can occur in this than diffusive layering. The additional constraint on Tz,c experiment at places where Tz is lower than Tz,c in the could eliminate some places where diffusive layering is standard DDP run. Examination of the vertical distri- possible due to weak Tz, but the ®ngers predominate bution of diffusivities reveals that double diffusion largely because of the restoring boundary conditions and reaches deeper in the experiment with higher vertical the deep water formation at high latitudes, which pro- resolution than in the standard run. Thus, an increase duces cold, fresh deep waters. Thus, under such upper in vertical resolution can improve the distribution of boundary conditions, diffusive mixing plays a small role water masses, but the meridional overturning rate re- in the diapycnal diffusion. In this case, the unphysical mains almost unchanged. representation of diffusive layering by a simple d ϭ KS/ K Ͼ 1 in experiment GHD does not affect the result T 5. Summary too much. However, if we use a virtual salt ¯ux condition or An investigation of the effects of double-diffusive natural boundary condition (Huang 1993), the situation mixing on the general circulation has been carried out should be different. Under the so-called mixed boundary using a more realistic parameterization than that used conditions, multiple equilibria exist. Convection could by Gargett and Holloway (1992), who applied a uni- happen near the equator rather than in the polar region versal nonunity ratio between the vertical diffusivities in the hemisphere model, or sinking near one pole while of heat and salt. The parameterization has three distinct upwelling near another pole in a global model, thus the features compared with the conventional constant eddy conditions favorable for diffusive layering are not af- diffusivity assumption: 1) The vertical eddy diffusiv- fected in places where no convection occurs, and then ities of temperature (KT ) and salinity (KS ) are different. it may be essential for the diapycnal mixing processes. For salt ®nger mixing, KS Ͼ KT , and vice versa for diffusive layering. 2) Both KT and KS increase when the local density ratio R ϭ ␣T /␤S is close to 1, and d. Sensitivity to the vertical resolution ␳ z z there is a cutoff in the R␳ dependence. 3) A constant The strength of double-diffusion activity depends on diapycnal diffusivity applies for non-double-diffusive the density ratio, R␳ ϭ ␣Tz/␤Sz, which involves the turbulent mixing. When R␳ Ͼ 3.0 or R␳ Ͻ 0.3, the vertical gradient of temperature and salinity. Vertical diapycnal diffusivity is dominated by turbulent mixing temperature and salinity gradients may be poorly sim- rather than double diffusion. This conservative imple- ulated in low-vertical resolution models. In addition, the mentation of a double-diffusive parameterization leads vertical layer scale of staircases observed is smaller than to a decrease of the downward buoyancy diffusion. the vertical grid scale (which produces the necessity for Three experiments have been carried out using the a parameterization). Thus, a natural question is how GFDL MOM2 code based on standard relaxation sensitive is the parameterization to the vertical resolu- boundary conditions for the surface temperature and tion. salinity. The ®rst run applied the above parameteriza- We ran an experiment with a doubled vertical reso- tion (DDP), the second run used the GH92 parame- lution (compared to run DDP) and kept all other pa- terization (GHD), and the third run used a constant rameters the same. The meridional overturning cell in diapycnal diffusivity (CDD). Compared to run CDD,

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 603 the meridional overturning rate in run DDP is reduced redistribution of buoyancy, which produces pressure by 22% because of the reduced buoyancy ¯ux, and the gradients that drive vertical shears that act on the is- poleward heat transport is reduced by 8%. In compar- opycnal temperature and salinity gradients (Schmitt ison, the corresponding decrease in run GHD is 10% 1990). Such self-enhancing effects are well recognized and 4% respectively. The density ratio distribution in in ®nescale thermohaline intrusions, but our results in- run DDP and GHD is more favorable for double dif- dicate that they operate on larger vertical scales as well. fusion than in run CDD, indicating that changes in These experiments with the ``standard'' relaxation buoyancy ¯uxes affected the pressure ®eld in a way boundary conditions are intended to presage an ex- that enhances the differential advection of heat and salt. amination of double-diffusive effects in circulation Note that changes in the density ratio distribution in models with mixed boundary conditions. It is our in- run DDP are larger than in run GHD. Finally, deep tention to evaluate the impact of double diffusion on temperatures and salinities become higher in runs DDP the stability of GCMs with a ¯ux, rather than restoring, and GHD than in run CDD, and changes in run DDP boundary condition for the salinity. Such models typ- are more than twice those in run GHD. ically display complex oscillatory behavior on a va- Results showed that the western boundary region is riety of timescales, and we anticipate a sensitivity of the most favorable place for double diffusion due to the circulation to the double-diffusive mixing in certain the differential advection of water masses by strong parameter regimes. However, even the present results, surface and deep western boundary currents. This is which reveal a reduction in the meridional overturning consistent with the known occurrence of salt ®ngers circulation, should motivate work toward an improved in the western tropical Atlantic (Schmitt et al. 1987) parameterization of the double-diffusive mixing pro- and their inferred role in water mass changes in the cesses. Caribbean Sea and Gulf of Mexico (Schmitz et al. 1993). Acknowledgments. Detailed comments from two re- Sensitivity experiments show that the values of K*, viewers helped us to improve the manuscript substan-

Rc , and n can affect the ratio of mixing due to double tially. J.Z. and R.W.S. were supported by a grant from diffusion over the non-double-diffusive background the Ocean Sciences Division of the National Science mixing. When K*orRc are larger or n is smaller, mix- Foundation, OCE94-15589. R.X.H was supported by ing due to double diffusion intensi®es. As a result, NSF through Grant OCE93-00706. diapycnal diffusion of buoyancy is reduced, and the thermohaline circulation rate declines. The additional REFERENCES constraint of Tz,c on Tz can affect the activity of double diffusion in the deep ocean and subpolar region, but Bryan, F., 1987: On the parameter sensitivity of primitive equation it has no signi®cant effect on the meridional overturn- ocean general circulation model. J. Phys. Oceanogr., 17, 970± 985. ing rate. Experiments with salt ®ngering only and dif- Bryan, K., 1969: A numerical method for the study of the circulation fusive layering only indicate that for a single-hemi- of the world ocean. J. Comput. Phys., 4, 347±376. sphere basin under relaxation conditions for both tem- Cummins, P. F., G. Holloway, and A. E. Gargett, 1990: Sensitivity perature and salinity, salt ®ngering is dominant while of the GFDL ocean general circulation model to a parameter- ization of vertical diffusion. J. Phys. Oceanogr., 20, 817±830. diffusive layering activity is barely noticeable. How- Danabasoglu, G., and J. C. McWilliams, 1995: Sensitivity of the ever, diffusive layering could play a more important global ocean circulation model to parameterizations of mesoscale role under ``mixed'' type upper boundary conditions tracer transports. J. Climate, 8, 2967±2987. and in models that include a polar ocean. Higher ver- Federov, K. N., 1988: Layer thicknesses and effective diffusivities tical resolution appeared to improve the distribution of in ``diffusive'' thermohaline convection in the ocean. Small- Scale Turbulence and Mixing in the Ocean, Vol. 46, J. Nihoul water masses, but the strength of the thermohaline cir- and B. Jamart, Eds., Elsevier, 471±480. culation remained about the same. From the limited Gargett, A. E., and R. W. Schmitt, 1982: Observations of salt ®ngers observational data available, the parameterization pro- in the central waters of the eastern North Paci®c. J. Geophys. posed in section 2 seems a reasonable ®rst step to test Res., 87(C10), 8017±8029. , and G. Holloway, 1992: Sensitivity of the GFDL ocean model the effects of double diffusion. Although we consider to different diffusivities for heat and salt. J. Phys. Oceanogr., our parameterization to be rather conservative, its im- 22, 1158±1177. plementation led to signi®cant changes in the circu- , and B. Ferron, 1996: The effects of differential vertical dif- lation. fusion of T and S in a box model of thermohaline circulation. One of the more surprising results of this work was J. Mar. Res., 54, 827±866. Garrett, C. J. R., and W. H. Munk, 1972: Oceanic mixing by breaking the ®nding that the introduction of double diffusion internal waves. Deep-Sea Res., 19, 823±832. led to even more favorable conditions for its occur- Gent, P. R., and J. C. McWilliams, 1990: Isopycnal mixing in ocean rence. Whereas in a one-dimensional simulation circulation models. J. Phys. Oceanogr., 20, 150±155. (Schmitt 1981) double diffusion operates to drive R , J. Willebrand, T. J. McDougall, and J. C. McWilliams, 1995: ␳ Parameterizing eddy-induced tracer transports in ocean circu- away from one, its application in this general circu- lation models. J. Phys. Oceanogr., 25, 463±474. lation model tended to generate even more areas with Gregg, M. C., 1989: Scaling turbulent dissipation in the thermocline. low density ratio. This is due to the double-diffusive J. Geophys. Res., 94, 9686±9698.

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC 604 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 28

, and T. B. Sanford, 1987: Shear and turbulence in the ther- perature and salinity staircases in the northwestern Weddell Sea. mohaline staircases. Deep-Sea Res., 34, 1689±1696. J. Phys. Oceanogr., 20, 295±306. Hamilton, J. M., N. S. Oakey, and D. E. Kelley, 1993: Salt ®nger Munk, W. H., 1966: Abyssal recipes. Deep-Sea Res., 13, 707±730. signatures in microstructure measurements. J. Geophys. Res., Neal, V. T., S. Neshyba, and W. Denner, 1969: Thermal strati®cation 98(C2), 2453±2460. in the Arctic Ocean. Science, 166, 373±374. Hellerman, S., and M. Rosenstein, 1983: Normal monthly stress over Neshyba, S., V. T. Neal, and W. W. Denner, 1971: Temperature and the world ocean with error estimates. J. Phys. Oceanogr., 13, conductivity measurements under Ice Island T-3. J. Geophys. 1093±1104. Res., 76, 8107±8120. Hogg, N., P. Biscaye, W. Gardner, and W. J. Schmitz, 1982: On the Pacanowski, R. C., 1995: MOM2 documentation, user's guide and transport and modi®cation of Antarctic Bottom Water in the reference manual. Tech. Rep. 3, GFDL Ocean Group Technical Vema Channel. J. Mar. Res., 40 (Suppl.), 231±263. Report 232 pp. [Available from online at http://www.gfdl.gov/ Huang, R. X., 1993: Real freshwater ¯ux as a natural boundary con- ϳkd/momwebpages/momwww.html.] dition for the salinity balance and thermohaline circulation Padman, L., and T. M. Dillon, 1987: Vertical ¯uxes through the Beau- forced by evaporation and precipitation. J. Phys. Oceanogr., 23, fort Sea thermohaline staircase. J. Geophys. Res., 92, 10 799± 2428±2446. 10 806. , 1995: Advances in Theories of Wind-Driven and Thermohaline , and , 1988: On the horizontal extent of the Canada Basin Circulation, a Lecture Series. University of Hawaii at Manoa, thermohaline steps. J. Phys. Oceanogr., 18, 1458±1462. SOEST, 213 pp. Polzin, K., 1996: Statistics of the Richardson Number: Mixing models , and J. Yang, 1996: Deep-water upwelling in the frictional west- and ®nestructure. J. Phys. Oceanogr., 26, 1409±1425. ern boundary layer. J. Phys. Oceanogr., 26, 2243±2250. , J. M. Toole, and R. W. Schmitt, 1995: Finescale parameteri- Huppert, H. E., 1971: On the stability of a series of double-diffusive zations of turbulent dissipation. J. Phys. Oceanogr., 25, 306± layers. Deep-Sea Res., 18, 1005±1021. 328. Ingham, M. C., 1966: The salinity extrema of the world ocean. Ph.D. , , J. R. Ledwell, and R. W. Schmitt, 1997: Spatial variability thesis, Oregon State University, 63 pp. [Available from Uni- of turbulent mixing in the abyssal ocean. Science, 276, 93±96. versity Micro®lms Inc., 300 North Zeeb Road, Ann Arbor, MI Redi, M. H., 1982: Oceanic isopycnal mixing by coordinate rotation. 48106-1346.] J. Phys. Oceanogr., 12, 1154±1158. Kelley, D. E., 1984: Effective diffusivities within oceanic thermo- Robertson, R., L. Padman, and M. D. Levine, 1995: Finestructure, haline staircases. J. Geophys. Res., 89, 10 484±10 488. microstructure, and vertical mixing processes in the upper ocean , 1990: Fluxes through diffusive staircases: A new formulation. in the western Weddel Sea. J. Geophys. Res., 100, 18 517±18 535. J. Geophys. Res., 95, 3365±3371. Ruddick, B., 1983: A practical indicator of the stability of the water Kunze, E., 1990: The evolution of salt ®ngers in inertial wave shear. column to double-diffusive activity. Deep-Sea Res., 30, 1105± J. Mar. Res., 48, 471±504. 1107. Ledwell, J. R., A. J. Watson, and C. S. Law, 1993: Evidence for slow , 1992: Intrusive mixing in a Mediterranean salt lens: Intrusion mixing across the pycnocline from an open-ocean tracer-release slopes and dynamical mechanisms. J. Phys. Oceanogr., 22, experiment. Nature, 364, 701±703. 1274±1285. Levitus, S., 1982: Climatological Atlas of the World Ocean. NOAA , and J. S. Turner, 1979: The vertical length scale of double- Prof. Paper No. 13. Govt. Printing Of®ce, Washington, DC, 173 diffusive intrusions. Deep-Sea Res., 26A, 23±40. pp. , and L. Zhang, 1989: The mythical thermohaline oscillator? J. Lueck, R., 1987: Microstructure measurements in a thermohaline Mar. Res., 47, 717±746. staircase. Deep-Sea Res., 34, 1677±1688. Schmitt, R. W., 1979: The growth rate of super-critical salt ®ngers. Mack, S. A., 1985: Two-dimensional measurements of ocean micro- Deep-Sea Res., 26A, 23±40. structure: The role of double diffusion. J. Phys. Oceanogr., 15, , 1981: Form of temperature±salinity relationship in the central 1581±1604. water: Evidence for double-diffusive mixing. J. Phys. Ocean- , 1989: Towed chain measurement of ocean microstructure. J. ogr., 11, 1015±1026. Phys. Oceanogr., 19, 1108±1129. , 1988: Mixing in a thermohaline staircase. Small-Scale Tur- , and H. C. Schoeberlein, 1993: Discriminating salt ®ngering bulence and Mixing in the Ocean, Vol. 46, Elsevier Oceanogr. from turbulence-induced microstructure: Analysis of towed tem- Ser., J. Nihoul and B. Jamart, Eds., Elsevier, 435±452. perature±conductivity chain data. J. Phys. Oceanogr., 23, 2073± , 1990: On the density ratio balance in the Central Water. J. Phys. 2106. Oceanogr., 20, 900±906. Marmorino, G. O., W. K. Brown, and W. D. Morris, 1987: Two- , 1994: Double diffusion in oceanography. Annu. Rev. Fluid dimensional temperature structure in the C-SALT thermohaline Mech., 26, 255±85. staircase. Deep-Sea Res., 34, 1667±1675. , H. Perkins, J. D. Boyd, and M. C. Stalcup, 1987: C-SALT: An Marotzke, J., 1997: Boundary mixing and the dynamics of three- investigation of the thermohaline staircase in the western tropical dimensional thermohaline circulation. J. Phys. Oceanogr., 27, North Atlantic. Deep-Sea Res., 34, 1697±1704. 1713±1728. Schmitz, W. J., J. R. Luyten, and R. W. Schmitt, 1993: On the Florida Martinson, D. G., 1990: Evolution of the Southern Ocean winter current T/S envelope. Bull. Mar. Sci., 53, 1048±1065. mixed layer and sea ice: open ocean deep water formation and Shen, C., and R. W. Schmitt, 1995: The wavenumber spectrum of ventilation. J. Geophys. Res., 95, 11 641±11 654. salt ®ngers. Double-Diffusive Convection, A. Brandt and H. Fer- McDougall, T. J., 1983: Greenland Sea Bottom Water formation: A nando, Eds., Vol. 94, Amer. Geophys. Union, 305±312. balance between advection and double-diffusion. Deep-Sea Res., Shirtcliffe, T. G. L., 1967: Thermosolutal convection: Observation of 30, 1109±1118. an overstable mode. Nature, 213, 489±490. , and J. R. Taylor, 1984: Flux measurements across a ®nger Stern, M. E., 1960: The ``salt fountain'' and thermohaline convection. interface at low values of the stability ratio. J. Mar. Res., 42, Tellus, 12, 172±175. 1±14. , 1975: Ocean Circulation Physics. Academic Press, 246 pp. , and J. A. Whitehead Jr., 1984: Estimates of the relative roles Stommel, H., and A. B. Arons, 1960: On the abyssal circulation of of diapycnal, isopycnal and double-diffusive mixing in Antarctic the world ocean. Part II: An idealized model of the circulation Bottom Water in the North Atlantic. J. Geophys. Res., 89, pattern and amplitude in oceanic basins. Deep-Sea Res., 6, 339± 10 479±10 483. 343. Muench, R. D., H. J. S. Fernando, and G. R. Stegen, 1990: Tem- Taylor, J., 1991: Laboratory experiments on the formation of salt

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC APRIL 1998 ZHANG ET AL. 605

®ngers after the decay of turbulence. J. Geophys. Res., 96(C7), , and T. M. C. Hughes, 1992: Stability and variability of the 12 497±12 510. thermohaline circulation and its link to climate. Trends Phys. Toole, J. M., K. L. Polzin, and R. W. Schmitt, 1994: Estimates of Oceanogr., 1, 15±70. diapycnal mixing in the abyssal ocean. Science, 264, 1120±1123. Yin, F. L., and I. Y. Fung, 1991: Net diffusivity in ocean general Turner, J. S., 1973: Buoyancy Effects in Fluids. Cambridge University circulation models with nonuniform grids. J. Geophys. Res., 96, Press, 367 pp. , 1978: Double-diffusive intrusions into a density gradient. J. 10 773±10 776. Geophys. Res., 83, 2887±2901. Zhang, J., R. W. Schmitt, and R. X. Huang, 1997: The relative in- Weaver, A. J., and E. S. Sarachik, 1990: On the importance of vertical ¯uence of diapycnal mixing and hydrologic forcing on the sta- resolution in certain ocean general circulation models. J. Phys. bility of the thermohaline circulation. J. Phys. Oceanogr., sub- Oceanogr., 20, 600±609. mitted.

Unauthenticated | Downloaded 09/23/21 07:45 PM UTC