Embryonic ζ- and ε-globin subunits assemble with each other and with adult α- and β-globin subunits into hemoglobin heterotetramers in both primitive and definitive erythrocytes. The properties of these hemoglobins—Hbs Gower-1 (ζ2ε2), Gower-2 (α2ε2), and Portland-2 (ζ2β2)—have been incompletely described as they are difficult to obtain in quantity from either primary human tissue or conventional expression systems. The generation of complex transgenic–knockout mice that express these hemoglobins at levels between 24% and 70% is described, as are efficient methods for their purification from mouse hemolysates. Key physiological characteristics—including P50, Hill coefficient, Bohr effect, and affinity for 2,3-BPG—were established for each of the 3 human hemoglobins. The stability of each hemoglobin in the face of mechanical, thermal, and chemical stresses was also determined. Analyses indicate that the ζ-for-α exchange distinguishing Hb Portland-2 and Hb A alters hemoglobin O2-transport capacity by increasing its P50 and decreasing its Bohr effect. By comparison, the ε-for-β exchange distinguishing Hb Gower-2 and Hb A has little impact on these same functional parameters. Hb Gower-1, assembled entirely from embryonic subunits, displays an elevated P50 level, a reduced Bohr effect, and increased 2,3-BPG binding compared to Hb A. The data support the hypothesis that Hb Gower-2, assembled from reactivated ε globin in individuals with defined hemoglobinopathies and thalassemias, would serve as a physiologically acceptable substitute for deficient or dysfunctional Hb A. In addition, the unexpected properties of Hb Gower-1 call into question a common hypothesis for its primary role in embryonic development.

Physiologically meaningful human hemoglobins assemble from 2 α-like and 2 β-like globin subunits. The 3 genes that encode α-like globins (5′-ζ-α2-α1-3′) and the 5 genes that encode β-like globins (5′-ε-Gγ-Aγ-δ-β-3′) are expressed in a developmental sequence that parallels their structural arrangement.1,2 Coordinated switching of the α-like and β-like genes results in the high-level expression of Hb Gower-1 (ζ2ε2), Hb F (α2γ2), and Hb A (α2β2) during the embryonic, fetal, and adult developmental stages, respectively. Other structurally defined hemoglobins—Gower-2 (α2ε2), Portland-1 (ζ2γ2), and Portland-2 (ζ2β2)—are expressed at relatively low levels in primitive and definitive erythroid cells, primarily during embryonic and early fetal development.1-8 Their heterotetrameric structures predict that each of these semi-embryonic hemoglobins will display properties compatible with human physiology.

Unlike Hb A (α2β2) and Hb F (α2γ2), the properties of semi-embryonic Hbs Gower-2 (α2ε2) and Portland-2 (ζ2β2), as well as fully embryonic Hb Gower-1 (ζ2ε2), remain largely undefined.3-8 Intact erythrocytes from 35-mm crown–rump embryos (approximately 7-week gestation), containing a complex mixture of embryonic, semi-embryonic, and fetal hemoglobins, have been shown to bind O2 strongly.9 However, it is difficult to purify individual hemoglobins from primitive erythroid cells because they are produced in low numbers during an early and relatively brief developmental window4,5 and because they contain a highly heterogeneous population of hemoglobin heterotetramers.4,5,9 Recently, a yeast expression system was developed to generate Gower hemoglobins for in vitro analysis.10 As with bacteria, yeast culture systems express relatively low levels of functional hemoglobin, which must be rigorously purified from incompletely processed globins and from inaccurately assembled heterotetramers.11-13 Nevertheless, yeast-expressed Gower hemoglobins appear to exhibit high O2affinities when studied under defined conditions.10 14 A more comprehensive functional and structural characterization of Gower and other low-abundance hemoglobins would be facilitated by the generation of a system that expresses high levels of fully functional heterotetramers.

The successful high-level expression of fully processed and fully assembled human Hbs A, F, S, and C in complex transgenic–knockout mice15-17 suggested that a similar strategy might be used to generate large quantities of less common human hemoglobins for in vitro analysis. A key step in this process was the generation of adult mice expressing high levels of human (h) embryonic ζ- and ε-globins in their definitive erythrocytes.18 These mice were subsequently used to generate lines expressing hybrid mouse–human semi-embryonic hemoglobins at 100% levels.19Comprehensive in vitro and in vivo evaluation demonstrated that hemoglobins assembled from mα- and hε-globin subunits (mα22) displayed O2-binding properties similar to those of control Hb mα22. This result strengthened the hypothesis that physiologically important characteristics of fully human Hb Gower-2 (α2ε2) might be similar to those of Hb A (α2β2).19 In contrast, substantial differences in the properties of Hbs hζ22 and hα22 suggested that the physiological characteristics of fully human Hb Portland-2 (ζ2β2) and Hb A might differ in several important respects.19 In addition to providing an estimate of the biochemical and physiological properties of the semi-embryonic hemoglobins, these studies also indicated the potential value of human embryonic globin subunits as substitutes for adult globin subunits in individuals with defined thalassemias or hemoglobinopathies.

The current study extends our previous work by assessing key biochemical and physiological properties of fully humansemi-embryonic and embryonic hemoglobins purified from complex transgenic–knockout mice. We describe a strategy for generating mice expressing high levels of human Hbs Gower-1 (ζ2ε2), Gower-2 (α2ε2), and Portland-2 (ζ2β2), as well as specific methods for their rapid and efficient purification. The key biochemical characteristics of each hemoglobin are subsequently determined, including their O2 affinities, subunit cooperativities, and changes in O2 affinity in response to allosteric modifiers and variations in ambient pH. We also assess the stability of each of these hemoglobins in response to defined mechanical, chemical, and thermal stresses. Based on the data, we speculate on the evolutionary basis for hemoglobin switching and the potential value of these poorly understood hemoglobins to patients with congenital α- and β-globin chain defects.

Transgenic and knockout mice

The generation and characterization of transgenic mice expressing high levels of human α, β, ζ, and ε globins have previously been described.18-21 Mice with heterozygous knockout of their endogenous α-globin genes (genotype mα+/−) or β-globin genes (genotype mβ+/−) were generously provided by Y. W. Kan and Judy Chang (University of California, San Francisco)22 and O. Smithies (University of North Carolina, Chapel Hill),23respectively. All mouse husbandry and experimentation was performed using protocols approved by the IACUC of the University of Pennsylvania.

Hemoglobin purification

Whole blood was collected from decapitated mice in 200 μL phosphate-buffered saline (PBS)–heparin (20 U/mL) or PBS–EDTA (27 mM), and the hemoglobins were promptly converted to the carbonmonoxy form by bubbling the sample with CO. Erythrocytes were subsequently washed twice with excess PBS–EDTA (2.7 mM), and the cell pellets were stored in aliquots at −80°C. Lysate was prepared in approximately 3-fold excess buffer A (see below) and clarified by ultracentrifugation at 20°C in a TLA-100 rotor at 40 000 rpm for 20 minutes (Beckman, Fullerton, CA). Hemolysates were fractionated over an SP/H 4.5 × 100 Poros column (PerSeptive Biosystems, Foster City, CA) using a BioCAD Sprint perfusion chromatography system (Framingham, MA). Hb Gower-1 was purified using buffer A (40 mM Bis-Tris, 5 mM EDTA, pH 6.5) and buffer B (buffer A + 200 mM NaCl) at 2 mL/min using a linear 10% to 60% buffer B gradient. Hb Portland-2 was similarly purified using a 10% to 40% buffer B gradient. To purify Hb Gower-2, buffers were adjusted to pH 6.8, and a nonlinear 30% to 50% buffer B gradient was used. Fractions collected in 96-well microtiter plates were analyzed at A540 on a SpectraMAX plate reader (Molecular Devices, Sunnyvale, CA), pooled, and concentrated at 4°C over a Centricon YM-10 filter (Millipore, Bedford, MA). Hb A, prepared from human hemolysate, was used as a control in all experiments.

Electrophoretic analysis

The identity and purity of each hemoglobin preparation was verified by denaturing Triton–acid–urea24,25 and nondenaturing cellulose acetate electrophoresis19 using methods recommended by the manufacturer (Helena Laboratories, Beaumont, TX).

Oxygen equilibrium curves

Purified CO-hemoglobins were resuspended to a final concentration of approximately 7.5 μM in P50 buffer (50 mM Bis-Tris, pH 7.4, 100 mM NaCl, 5 mM EDTA) and converted to the oxy form by photolysis under 100% O2 using an ice water–cooled rotary condenser.26 Conversion to the oxyhemoglobin form was judged complete by an A540:A576 ratio of less than 0.95. Oxygen equilibrium curves (OECs) were subsequently determined on a HEMOX analyzer (TCS, Southampton, PA) at 20°C. Studies of 2,3-bisphosphoglycerate binding (2,3-BPG; Sigma, St Louis, MO) were carried out in P50 buffer (pH 7.4), whereas Bohr effect studies were carried out in P50 buffer adjusted to defined pH values.

Stability determinations

Mechanical.

Using a modified version of a previously described method,27 Hbs were diluted to approximately 13 μM with 10 mM potassium phosphate buffer (pH 8.0) and converted to the oxyhemoglobin form (see above). Aliquots (2 mL) were shaken for defined intervals at a setting of 2000 on a Maxi-Mix III type 65800 shaker (Thermolyne, Dubuque, IA), and denatured hemoglobins were precipitated by a 5-minute desktop spin. The soluble hemoglobin was determined by A542 spectrophotometry of the supernatant.

Chemical.

Purified hemoglobins were diluted to approximately 0.1 mM in buffer (0.1 mM Tris, pH 7.4) and converted to the oxyhemoglobin form as described above. Aliquots diluted 10-fold in prewarmed Tris buffer containing 17% (vol/vol) isopropanol were incubated at 37°C for 5 minutes.28 Precipitated hemoglobins were clarified by desktop centrifugation, and the A542 of the supernatant was determined.

Thermal.

Purified hemoglobins were diluted to approximately 50 μM in buffer (0.1 mM Tris, pH 7.4) and converted to the oxyhemoglobin form as described above. Test and control aliquots were incubated for 2 hours at 50°C and 4°C, respectively, and spun for 10 minutes on a desktop centrifuge.29 The supernatant was diluted 10-fold with developer solution [11.9 mM NaHCO3, 0.77 mM KCN, 0.61 mM K3Fe(CN)6], insoluble hemoglobins were precipitated by desktop centrifugation, and the A540 of the supernatant was determined.

Generation of adult mice expressing high levels of human embryonic and semi-embryonic hemoglobins

The construction of transgenes and the generation of mice expressing hζ, hα, hε, and hβ globins in definitive erythrocytes has previously been described.18-20 The high-level, developmental-stage inappropriate expression of transgenic hζ and hε globins was achieved by linking their encoding genes to transcriptional control elements from the hα and hβ globin genes, respectively (Figure 1A).18Full-length genes encoding hα and hβ globins, containing their native transcriptional control elements, were anticipated to be expressed at high levels in adult erythrocytes and consequently were not modified.19,20 Each transgene was linked to a micro β-locus control region to insure its high-level, integration position-independent expression.18-20,30 Single lines expressing high levels of each transgenic globin were identified by phenotypical screening of hemolysates using denaturing globin electrophoresis.24 25 These lines were used in the experiments described in the current work.

Fig. 1.

Generation of transgenic mice expressing high levels of human embryonic and semi-embryonic hemoglobins.

(A) Structures of human transgenes. Exons are depicted as open boxes, with the positions of the translational initiation and termination codons indicated by tick marks. Globin-gene origins of promoter and enhancer elements (cross-hatched) are also indicated. All transgenes were linked to an identical micro β-LCR cassette (shaded).30 The common name for each transgene and the human globin it expresses are indicated to the left and right of the diagram, respectively. (B) Mating strategy for generating mice expressing human Hb Gower-1 (ζ2ε2). Partial globin genotypes of selected mice from 5 generations (F1-F5) are depicted. The strategy facilitates the generation of mice expressing high levels of human Hb ζ2ε2 from progenitors expressing hζ or hε globin or containing heterozygous deletion of their endogenous mα or mβ globin genes. A similar strategy was used to generate mice expressing high levels of human Hbs Gower-2 (α2ε2) and Portland-2 (ζ2β2). m, mouse; h, human; +/+, homozygous; +/−, heterozygous; −/−, nullizygous.

Fig. 1.

Generation of transgenic mice expressing high levels of human embryonic and semi-embryonic hemoglobins.

(A) Structures of human transgenes. Exons are depicted as open boxes, with the positions of the translational initiation and termination codons indicated by tick marks. Globin-gene origins of promoter and enhancer elements (cross-hatched) are also indicated. All transgenes were linked to an identical micro β-LCR cassette (shaded).30 The common name for each transgene and the human globin it expresses are indicated to the left and right of the diagram, respectively. (B) Mating strategy for generating mice expressing human Hb Gower-1 (ζ2ε2). Partial globin genotypes of selected mice from 5 generations (F1-F5) are depicted. The strategy facilitates the generation of mice expressing high levels of human Hb ζ2ε2 from progenitors expressing hζ or hε globin or containing heterozygous deletion of their endogenous mα or mβ globin genes. A similar strategy was used to generate mice expressing high levels of human Hbs Gower-2 (α2ε2) and Portland-2 (ζ2β2). m, mouse; h, human; +/+, homozygous; +/−, heterozygous; −/−, nullizygous.

Close modal

A breeding strategy was designed to generate mice expressing high levels of human embryonic or semi-embryonic hemoglobins with minimal mouse globin background (Figure 1B). We had previously noted that the expression of each of the 4 human globin transgenes increased substantially in mice carrying one or more knockout mutations of the related endogenous adult globin gene homologue.18,19 The level of hζ and hα induction was sufficient to rescue the viability of mice with homozygous-lethal deletions of their endogenous mα-globin genes,18,31,32 whereas the viability of mice with homozygous-lethal deletions of the mβ-globin genes could be rescued by the expression of either transgenic hε or hβ globins.18,31 32 We reasoned that the assembly of hemoglobin heterotetramers from transgenic human α-like and β-like globins would be similarly enhanced in mice carrying both mα- and mβ-globin knockout alleles.

Design of methods to purify human hemoglobins from transgenic hemolysates

A combination of genetic and biochemical strategies was used to facilitate the preparation of human Hbs from transgenic mice. We screened more than 125, 335, and 89 candidate pups expressing Hbs Gower-1, Gower-2, and Portland-2, respectively, without identifying any mα−/−/mβ−/− mice expressing 100% of the desired human hemoglobins (data not shown). On the other hand, a substantial proportion of these pups displayed either mα+/−/mβ−/− or mα+/+/mβ−/− genotypes (more than 25 pups expressing each hemoglobin; data not shown). Hbs Gower-1 (ζ2ε2), Gower-2 (α2ε2), and Portland-2 (ζ2β2) were expressed in the these complex transgenic–knockout mice as 37%, 24%, and approximately 70% of total hemoglobin, respectively, corresponding to the expression of human hemoglobin in the range of approximately 20 to 80 mg/mouse (data not shown). These high levels of expression facilitated the task of hemoglobin purification, as did the fact that mα+/−/mβ−/− or mα+/+/mβ−/− mice each assembled only a single contaminant hemoglobin species (mα22 or mα22).

A method was subsequently established for isolating each of the desired human hemoglobin heterotetramers using cation-exchange chromatography. Human Hbs Gower-1 (ζ2ε2), Gower-2 (α2ε2), and Portland-2 (ζ2β2) were purified from contaminant hybrid Hbs mα22, mα22, and mα22, respectively, in single-step processes using NaCl gradients at defined pH levels (Figure2, and data not shown). The identities and purities of the eluted human hemoglobins were subsequently verified by nondenaturing19 and denaturing24 25electrophoretic methods (Figure 2 and data not shown). The large quantities of high-purity human embryonic and semi-embryonic hemoglobins efficiently prepared by this method were sufficient to permit their detailed physiological and biochemical evaluation and to provide substantial banked product for future functional and structural studies.

Fig. 2.

Purification of human embryonic and semi-embryonic hemoglobins.

(A) Human Hb Gower-1 (ζ2ε2). (Left) Hemolysate prepared from adult mα+/−/mβ−/−/hζ/hε complex transgenic–knockout mice was resolved over a Poros SP/H column using a linear 20 to 120 mM NaCl gradient (pH 6.5). The positions of human Hb ζ2ε2 (peak 1) and contaminant hybrid mouse–human Hb mα22 (peak 2) are indicated (arbitrary A280 units). Eluate conductivity (in mS) is depicted by a gray line. (Right) Aliquots of unfractionated (U) lysate and eluate corresponding to peaks 1 and 2 were resolved by nondenaturing cellulose acetate electrophoresis. A control lane contains a mixture of human Hbs A, F, S, and C. The migration of constituent hemoglobins is indicated to the left, and gel polarity to the right. (B) Human Hb Gower-2 (α2ε2). Resolution of hemolysate from an adult mα+/−/mβ−/−/hα/hε complex transgenic–knockout mouse into human Hb α2ε2 (peak 2) and contaminant hybrid mouse/human Hb mα22 (peak 1) using a nonlinear 60- to 100-mM NaCl gradient (pH 6.8). (C) Human Hb Portland-2 (ζ2β2). Resolution of hemolysate from an adult mα+/−/mβ−/−/hζ/hβ complex transgenic–knockout mouse into human Hb ζ2β2 (peak 1) and contaminant hybrid mouse/human Hb mα22 (peak 2) using a linear 20- to 80-mM NaCl gradient (pH 6.5).

Fig. 2.

Purification of human embryonic and semi-embryonic hemoglobins.

(A) Human Hb Gower-1 (ζ2ε2). (Left) Hemolysate prepared from adult mα+/−/mβ−/−/hζ/hε complex transgenic–knockout mice was resolved over a Poros SP/H column using a linear 20 to 120 mM NaCl gradient (pH 6.5). The positions of human Hb ζ2ε2 (peak 1) and contaminant hybrid mouse–human Hb mα22 (peak 2) are indicated (arbitrary A280 units). Eluate conductivity (in mS) is depicted by a gray line. (Right) Aliquots of unfractionated (U) lysate and eluate corresponding to peaks 1 and 2 were resolved by nondenaturing cellulose acetate electrophoresis. A control lane contains a mixture of human Hbs A, F, S, and C. The migration of constituent hemoglobins is indicated to the left, and gel polarity to the right. (B) Human Hb Gower-2 (α2ε2). Resolution of hemolysate from an adult mα+/−/mβ−/−/hα/hε complex transgenic–knockout mouse into human Hb α2ε2 (peak 2) and contaminant hybrid mouse/human Hb mα22 (peak 1) using a nonlinear 60- to 100-mM NaCl gradient (pH 6.8). (C) Human Hb Portland-2 (ζ2β2). Resolution of hemolysate from an adult mα+/−/mβ−/−/hζ/hβ complex transgenic–knockout mouse into human Hb ζ2β2 (peak 1) and contaminant hybrid mouse/human Hb mα22 (peak 2) using a linear 20- to 80-mM NaCl gradient (pH 6.5).

Close modal

Embryonic and semi-embryonic hemoglobins exhibit elevated O2 affinities

The O2-binding affinities of human Hbs ζ2ε2, α2ε2, and ζ2β2 were determined on 3 or more occasions under standard conditions by HEMOX analysis (Figure3A-C, Table1).19 Each of the hemoglobins displayed a higher O2 affinity than control Hb A, whose P50 of 3.2 torr was highly reproducible. The P50 values of Hbs Gower-1 (ζ2ε2) and Portland-2 (ζ2β2) (1.4 and 1.9 torr, respectively) were approximately one-half the P50 for control Hb A. In contrast, Hb Gower-2 (α2ε2) exhibited a P50 of 2.7 torr, only marginally different from that of Hb A. These results were substantiated in independent experiments, using different temperature and buffer conditions, in which the relative P50 values of the 4 human hemoglobins were preserved (data not shown).14 Hill coefficients derived from OEC analyses indicated substantially reduced subunit cooperativity for Hbs Gower-1 (ζ2ε2) and Portland-2 (ζ2β2) (Hill n = 1.7 and 1.6, respectively) relative to control Hb A (Figure 3D-F; Table 1). In contrast, Hb Gower-2 (α2ε2) (n = 2.3) displayed subunit cooperativity much closer to our observed measure for Hb A (n = 2.9), which reproduced its accepted value (n = 2.8-3.0).1 These results indicate that the O2-binding properties of Hb α2β2 heterotetramers are not materially affected by a β-to-ε exchange (converting Hb A to Hb Gower-2), whereas an α-to-ζ exchange (converting Hb A to Hb Portland-2) has a more substantial impact on both O2 affinity and subunit cooperativity.

Fig. 3.

Oxygen equilibrium curves and Hill coefficients for human embryonic and semi-embryonic hemoglobins.

Oxygen equilibrium curves are displayed in panels A-C. (A) Human Hb Gower-1 (ζ2ε2). (B) Human Hb Gower-2 (α2ε2). (C) Human Hb Portland-2 (ζ2β2). OECs were established for affinity-purified human hemoglobins under standard assay conditions (“Materials and methods”). Representative curves (black) are displayed with an OEC from control human Hb A for reference (gray). P50 values derived from analyses of these curves are included in Table 1. Hill coefficients are displayed in panels D-F. (D) Human Hb Gower-1 (ζ2ε2, ⧫). (E) Human Hb Gower-2 (α2ε2, ●). (F) Human Hb Portland-2 (ζ2β2, ■). Hill plots constructed from OECs of human hemoglobins in panels A to C are illustrated along with control human Hb A (O). Hill coefficients derived from these curves are included in Table 1.

Fig. 3.

Oxygen equilibrium curves and Hill coefficients for human embryonic and semi-embryonic hemoglobins.

Oxygen equilibrium curves are displayed in panels A-C. (A) Human Hb Gower-1 (ζ2ε2). (B) Human Hb Gower-2 (α2ε2). (C) Human Hb Portland-2 (ζ2β2). OECs were established for affinity-purified human hemoglobins under standard assay conditions (“Materials and methods”). Representative curves (black) are displayed with an OEC from control human Hb A for reference (gray). P50 values derived from analyses of these curves are included in Table 1. Hill coefficients are displayed in panels D-F. (D) Human Hb Gower-1 (ζ2ε2, ⧫). (E) Human Hb Gower-2 (α2ε2, ●). (F) Human Hb Portland-2 (ζ2β2, ■). Hill plots constructed from OECs of human hemoglobins in panels A to C are illustrated along with control human Hb A (O). Hill coefficients derived from these curves are included in Table 1.

Close modal
Table 1.

Properties of human hemoglobins

HbStructureP50*HillBohr2,3-BPG1-153
Gower-1 ζ2ε2 1.4 ± 0.06 1.7 ± 0.24 −0.25 0.09 
Gower-2 α2ε2 2.7 ± 0.10 2.3 ± 0.02 −0.51 0.17 
Portland-2 ζ2β2 1.9 ± 0.17 1.6 ± 0.06 −0.10 0.30 
α2β2 3.2 ± 0.14 2.9 ± 0.36 −0.54 0.29 
HbStructureP50*HillBohr2,3-BPG1-153
Gower-1 ζ2ε2 1.4 ± 0.06 1.7 ± 0.24 −0.25 0.09 
Gower-2 α2ε2 2.7 ± 0.10 2.3 ± 0.02 −0.51 0.17 
Portland-2 ζ2β2 1.9 ± 0.17 1.6 ± 0.06 −0.10 0.30 
α2β2 3.2 ± 0.14 2.9 ± 0.36 −0.54 0.29 
*

torr;

log(Y/1 − Y)/log PO2;

Δlog P50/ΔpH in the alkaline region;

F1-153

apparent binding constant (mM) calculated from half-saturation point.

The P50 values of embryonic and semi-embryonic hemoglobins display disparate responses to changes in pH and [2,3-BPG]

The effect of pH on the O2-binding affinities of the 3 human embryonic and semi-embryonic hemoglobins was determined by measuring the P50 values for each in a series of buffers with defined pH levels ranging from 6.0 to 8.2 (Figure4A; Table 1). The 2 human hemoglobins containing ζ-globin subunits displayed attenuated Bohr effects [Hbs ζ2ε2 (−0.10 ΔlogP50/ΔPO2) and ζ2β2 (−0.25 ΔlogP50/ΔPO2)], whereas the Bohr effect of Hb Gower-1 (α2ε2) and control Hb A (α2β2) were nearly identical (−0.51 vs −0.54 ΔlogP50/ΔPO2, respectively). The apparent binding constant of 2,3-BPG for each hemoglobin was estimated by establishing the half-saturation point using buffers with defined concentrations of the allosteric modifier (Figure 4B; Table 1). 2,3-BPG appears to bind to Hb Portland-2 (ζ2β2) and control Hb A with equal affinity (0.30 mM and 0.29 mM, respectively) while binding to Hbs Gower-1 (ζ2ε2) and Gower-2 (α2ε2) with substantially higher avidity (0.09 mM and 0.17 mM, respectively). These results indicate that the inclusion of embryonic globin subunits may have an impact on the biochemical function of intact heterotetramers, permitting speculation on the evolutionary pressures favoring the conservation of globin gene switching.

Fig. 4.

Biochemical properties of human embryonic and semi-embryonic hemoglobins.

(A) Effect of pH on O2 binding (Bohr effect). The P50 of each hemoglobin was determined in standard buffers adjusted to defined pH levels. Bohr effect values (included in Table 1) were calculated from best-fit curves of values from the alkaline range. ζ2ε2 (⧫), α2ε2 (●), ζ2β2 (■), α2β2 (○). (B) Effect of allosteric modifiers on O2 binding. The P50 values of human hemoglobins were determined in standard buffers containing defined concentrations of 2,3-BPG. The affinity of each Hb for 2,3-BPG (indicated in Table 1) was derived from the half-saturation point of each curve. ζ2ε2 (⧫), α2ε2 (●), ζ2β2 (■), α2β2 (○).

Fig. 4.

Biochemical properties of human embryonic and semi-embryonic hemoglobins.

(A) Effect of pH on O2 binding (Bohr effect). The P50 of each hemoglobin was determined in standard buffers adjusted to defined pH levels. Bohr effect values (included in Table 1) were calculated from best-fit curves of values from the alkaline range. ζ2ε2 (⧫), α2ε2 (●), ζ2β2 (■), α2β2 (○). (B) Effect of allosteric modifiers on O2 binding. The P50 values of human hemoglobins were determined in standard buffers containing defined concentrations of 2,3-BPG. The affinity of each Hb for 2,3-BPG (indicated in Table 1) was derived from the half-saturation point of each curve. ζ2ε2 (⧫), α2ε2 (●), ζ2β2 (■), α2β2 (○).

Close modal

Human embryonic and semi-embryonic hemoglobins display different physical stabilities

Hbs Gower-1, Gower-2, and Portland-2 were assessed for their stability in the setting of defined mechanical,27chemical,28 and thermal stresses.29 Hbs A and S were evaluated in parallel as stable and unstable hemoglobin controls, respectively (Figure 5). The stabilities of the various hemoglobins by each of the 3 methods were in general agreement: Hbs Gower-1 (ζ2ε2) and Portland-2 (ζ2β2) were equally or less stable than Hb S, whereas the stability of Hb Gower-2 (α2ε2) was generally intermediate between the 2 control hemoglobins. Hence, the low stability of Hb Gower-1 appears to be well adapted to the short survival of primitive erythrocytes, whereas the higher stability of Hb Gower-2 may facilitate its expression in long-lived definitive erythrocytes.

Fig. 5.

Stabilities of human hemoglobins to defined perturbations.

(A) Mechanical stability. The stabilities of the human hemoglobins exposed to mechanical stress were determined as described in “Materials and methods.” The percentage soluble hemoglobin is plotted as a function of time. Early time points are displayed on an expanded scale (inset). (B) Chemical stability. The stabilities of the human hemoglobins exposed to 17% isopropanol are illustrated. Bars represent the average of duplicate determinations using independently prepared isopropanol solutions. Symbols are the same as in panel A. (C) Thermal stability. The relative stabilities of the human hemoglobins incubated at 50°C were determined and plotted as a function of precipitated hemoglobin. Symbols are the same as in panel A.

Fig. 5.

Stabilities of human hemoglobins to defined perturbations.

(A) Mechanical stability. The stabilities of the human hemoglobins exposed to mechanical stress were determined as described in “Materials and methods.” The percentage soluble hemoglobin is plotted as a function of time. Early time points are displayed on an expanded scale (inset). (B) Chemical stability. The stabilities of the human hemoglobins exposed to 17% isopropanol are illustrated. Bars represent the average of duplicate determinations using independently prepared isopropanol solutions. Symbols are the same as in panel A. (C) Thermal stability. The relative stabilities of the human hemoglobins incubated at 50°C were determined and plotted as a function of precipitated hemoglobin. Symbols are the same as in panel A.

Close modal

In spite of their developmental sobriquets, there is considerable overlap in the temporal expression of embryonic ζ and ε globins and adult α and β globins, respectively. Human α and β globins are expressed at moderate levels in developing embryos and fetuses,33-36 whereas varying amounts of embryonic ζ and ε globins and their encoding mRNAs can be detected in fetal erythrocytes and in adult-stage reticulocytes, respectively.37,38 The developmental integrity of embryonic globin gene expression is further compromised in patients with certain congenital genetic disorders. Low-level ζ-globin expression persists in adults heterozygous for the α-thalassemia –SEA deletion,39 whereas fetuses and infants with defined trisomies may express easily detectable levels of ε globin.40-43 The temporal overlap of embryonic and adult globin expression, particularly during intrauterine development, favors the assembly of Hbs Gower-2 and Portland-2. Remarkably few studies have directly addressed the biochemical and physiological properties of either of these hemoglobins or of the embryonic hemoglobin Gower-1, despite their assembly in normal primitive and definitive erythrocytes.

Although it is uncertain whether Hbs Gower-2 and Portland-2 play a defined role in human development or are simply incidental to developmental overlapping of globin gene expression, a full understanding of their properties, along with the properties of Hb Gower-1, would be valuable for several purposes. First, structure–function analyses of all 3 hemoglobins are likely to provide additional insight into the biochemistry of abundant, thoroughly studied hemoglobins such as Hbs A and F. Second, knowledge of its properties would provide a context for assessing the role of Hb Gower-1 in normal development and would help focus speculation on the evolutionary pressures favoring absolute phylogenic conservation of developmental globin gene switching. Third, a direct application of the analyses of Hbs Gower-2 and Portland-2 stems from the possibility that reactivation of the ζ and ε genes would produce globins that could substitute for deficient or abnormal α- or β-globin chains, respectively, in individuals with defined thalassemias and hemoglobinopathies.18 19 The implications of the current work on each of these 3 issues are considered below.

Many well-established studies have identified functionally crucial amino acid residues in the α- and β-globin chains. Our data generally confirm the role of these residues, but they also emphasize the importance of distant residues on specific hemoglobin functions through their likely influence on the high-order structure of the globin molecule. In general, Hb Gower-2 and Hb A function similarly, suggesting that many of the 36 amino acid differences between the β- and ε-globin chains (of 146 total residues) cluster in functionally silent domains (Table 1). The 2 hemoglobins exhibit similar O2-binding characteristics and display Hill coefficients consistent with normal hemoglobin function (Figures 3B, E), reflecting the conservation of 13 of 17 and 15 of 16 residues comprising the α1β1 and α1β2 interfaces, respectively.44,45Similarly, the robust Bohr effect common to Hb Gower-2 and Hb A (Figure4A) might be anticipated from the conservation of 82Lys, 143His, and 146His in both the ε- and β-globin chains.46-48Unexpectedly, Hb Gower-2 and Hb A displayed markedly different affinities for 2,3-BPG (Figure 4B) despite their identity at all 4 residues believed to mediate this property (1Val, 2His, 82Lys, and 143His).46,49 50 Although the half-saturation method provides only an estimate of the strength of this interaction, it is likely that one or more differences in amino acid content elsewhere in the ε- and β-globin chains alter the spatial arrangement, and hence the function, of the 4 conserved residues. A less likely explanation is that a fifth, unrecognized residue, differing between the β- and ε-globin chains, directly participates in 2,3-BPG binding. Crystallographic evaluation of Hb Gower-2, available in quantity from the transgenic–knockout mice, would be useful in evaluating these 2 possibilities.

In comparison to the β-like globin chains, the 57 (of 141) amino acid differences between the ζ- and α-globin chains appear to occupy functionally sensitive regions (Table 1). Hb Portland-2 and Hb A, differing only in the identity of their α-like chains, display substantially different P50 values and Hill coefficients (Figure 3C, F) despite sharing 14 of 16 and 14 of 15 residues at the α1β1 and α1β2 interfaces, respectively. Remarkably, 2 of the 3 interface substitutions are conservative, emphasizing the unpredictable functional effects of changes in high-order globin subunit structure resulting from spatially distant amino acid substitutions. Although both globins share a common 122His, the α1Val-ζ1Ser exchange would be expected to substantially blunt the Bohr effect exhibited by Hb Portland-2,46 51 a prediction that is experimentally observed (Figure 4A). A direct comparison of Hb Portland-2 function with its high-resolution structure, when available, may yield important new information relative to the function of human heterotetramers.

The observation that the properties of Hb Gower-1, comprising developmental stage-concordant ζ- and ε-globin subunits, are strikingly different from those of Hb A suggests a strong evolutionary basis for Hb switching (Figures 3A, 4A-B; Table 1). A widely held, yet largely unsubstantiated, explanation for this process suggests that embryonic and fetal hemoglobins have been evolutionarily selected to facilitate the trans-placental delivery of O2during intrauterine development. This hypothesis has been challenged by studies that do not detect the predicted elevated rates of fetal wastage in pregnant women with high-affinity hemoglobins.52,53 By demonstrating that the O2-binding properties of Hb Gower-1 differ from those of Hb F, our data support the latter dissenting opinion. In contrast to Hb F, which exhibits a normal O2 affinity (in the absence of 2,3-BPG), a relatively high Bohr effect, and poor 2,3-BPG binding, we have found that Hb Gower-1 displays a high O2 affinity, a reduced Bohr effect, and relatively tight 2,3-BPG binding. Were their developmental roles strictly limited to O2-transport, the O2-binding properties of Hbs Gower-1 and F might be anticipated to be more similar. An alternative explanation for this difference is that the evolutionary basis for developmental hemoglobin switching reflects properties of Hb Gower-1 that are independent of its gaseous-exchange function. This possibility is underscored by recent reports indicating that in addition to its O2-transporting function, Hb A may serve an important role in the regulation of vascular tone and blood flow.54 It is clear that the functional properties of Hb Gower-1, reflecting poorly understood evolutionary demands and manifesting as phylogenic conservation of hemoglobin switching, merit additional study.

We have previously proposed that Hbs Portland-2 (ζ2β2) and Gower-2 (α2ε2) might be suitable substitutes for Hb A in individuals with defined hemoglobinopathies and thalassemias.18,19 These 2 hemoglobins would assemble from existing adult globin chains in adults in whom the embryonic globin genes were reactivated. The functional identity between Hb Gower-2 (α2ε2) and Hb A (α2β2) indicates that the former hemoglobin would act in a physiologically valuable manner in definitive erythrocytes. Evidence that the ε- and γ-globin genes are independently regulated55-58 raises the possibility that different classes of agents, with different toxicities, might be used to separately reactivate expression from the 2 genes. This consideration is particularly important in patients with severe β-thalassemia determinants in whom γ-globin reactivation by current agents is toxicity-limited.59 60 The observation that Hb Gower-2 will serve a physiologically important role in individuals with any of several hemoglobinopathies and thalassemias should serve as an impetus to identify agents that might be useful in this regard.

We thank S. Krishnaswami and K. Adachi for access to protein purification and analytical instruments and Y. Yang for technical assistance. J.E.R. is the recipient of a Junior Faculty Scholar Award from the American Society of Hematology.

Supported by National Institutes of Health grant R01 HL61399 and past support from Cooley's Anemia Foundation.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 U.S.C. section 1734.

1
Bunn
HF
Forget
BG
Hemoglobin: molecular, genetic, and clinical aspects.
1986
WB Saunders
Philadelphia
2
Russell
JE
Liebhaber
SA
Molecular genetics of thalassemia.
Advances in Genome Biology.
Verma
RS
2
1993
283
353
JAI Press
Greenwich
3
Hecht
F
Motulsky
A
Lemire
R
Shepard
T
Predominance of human hemoglobin Gower-1 in early human embryonic development.
Science.
152
1966
91
92
4
Huehns
ER
Flynn
FV
Butler
EA
Beaven
GH
Two new haemoglobin variants in a very young embryo.
Nature.
189
1961
496
497
5
Huehns
E
Dance
N
Beaven
G
Hecht
F
Motulsky
A
Human embryonic hemoglobins.
Cold Spring Harb Symp Quant Biol.
29
1964
327
331
6
Huehns
E
Beaven
G
Developmental changes in human hemoglobins.
Clin Dev Med.
37
1971
175
203
7
Gale
R
Clegg
J
Huehns
E
Human embryonic hemoglobins Gower-1 and Gower-2.
Nature.
280
1979
162
164
8
Randhawa
Z
Jones
R
Lie-Injo
L
Human hemoglobin Portland II (ζ2β 2).
J Biol Chem.
259
1984
7325
7330
9
Huehns
ER
Farooqui
AM
Oxygen dissociation properties of human embryonic red cells.
Nature.
234
1975
335
337
10
Mould
RM
Hofmann
OM
Brittain
T
Production of human embryonic hemoglobin (Gower-II) in a yeast expression system.
Biochem J.
298
1994
619
622
11
Hoffman
S
Looker
D
Roehrich
J
et al
Expression of fully functional tetrameric human hemoglobin in Escherichia coli.
Proc Natl Acad Sci U S A.
87
1990
8521
8525
12
Shen
T-J
Ho
N
Simplaceanu
V
et al
Production of unmodified human adult hemoglobin in Escherichia coli.
Proc Natl Acad Sci U S A.
90
1993
8108
8112
13
Adachi
K
Konitzer
P
Lai
CH
Kim
J
Surrey
S
Oxygen-binding properties and other physical properties of human hemoglobin made in yeast.
Protein Eng.
5
1992
807
810
14
Hofmann
OM
Mould
R
Brittain
T
Allosteric modulation of oxygen binding to the three human embryonic hemoglobins.
Biochem J.
306
1995
367
370
15
Pastzy
C
Brion
CM
Manci
E
et al
Transgenic knockout mice with exclusively human sickle hemoglobin and sickle cell disease.
Science.
278
1997
876
879
16
Ryan
TM
Ciavatta
DJ
Townes
TM
Knockout-transgenic mouse model of sickle cell disease.
Science.
278
1997
873
876
17
Romero
J
Suzuka
S
Romero-Gonzalex
G
Nagel
R
Fabry
M
Knockout mouse red cells expressing human HbC or HbS and γ have high K:Cl cotransport activity in contrast to HbA mouse red cells
Program of the 24th Annual Meeting of the National Sickle Cell Disease Program, April 9-12
2000
; Philadelphia. Abstract 112.
18
Russell
JE
Liebhaber
SA
Reversal of lethal α and β thalassemias in mice by expression of human embryonic globins.
Blood.
92
1998
3057
3063
19
He
Z
Lian
L
Asakura
T
Russell
JE
Functional effects of replacing human α and β globins with their embryonic globin homologues in defined haemoglobin heterotetramers.
Br J Haematol.
109
2000
882
890
20
Liebhaber
SA
Wang
Z
Cash
F
Monks
B
Russell
JE
Developmental silencing of the embryonic ζ-globin gene: concerted action of the promoter and the 3′ flanking region combined with stage-specific silencing of the transcribed region.
Mol Cell Biol.
16
1996
2637
2646
21
Morales
J
Russell
JE
Liebhaber
SA
Destabilization of human alpha-globin mRNA by translation anti-termination is controlled during erythroid terminal differentiation and paralleled by phased shortening of the poly(A) tail.
J Biol Chem.
272
1997
6607
6613
22
Chang
J
Lu
RH
Xu
SM
et al
Inactivation of mouse α-globin gene by homologous recombination: mouse model of hemoglobin H disease.
Blood.
88
1996
1846
1851
23
Detloff
P
Lewis
J
John
S
et al
Deletion and replacement of the mouse adult β-globin genes by a “plug and socket” repeated targeting strategy.
Mol Cell Biol.
14
1994
6936
6943
24
Alter
BP
Gel electrophoretic separation of globin chains.
Prog Clin Biol Res.
60
1981
157
175
25
Rovera
G
Magarian
C
Borun
TW
Resolution of hemoglobin subunits by electrophoresis in acid urea polyacrylamide gels containing Triton X-100.
Anal Biochem.
85
1978
506
518
26
Shen
T
Ho
N
Zou
M
et al
Production of human normal adult and fetal hemoglobins in Escherichia coli.
Protein Eng.
10
1997
1085
1097
27
Asakura
T
Ohnishi
T
Friedman
S
Schwartz
E
Abnormal precipitation of oxyhemoglobin S by mechanical shaking.
Proc Natl Acad Sci U S A.
71
1974
1594
1598
28
Carrell
R
Kay
R
A simple method for the detection of unstable haemoglobins.
Br J Haematol.
23
1972
615
29
Schneiderman
LJ
Junga
IG
Fawley
DE
Effect of phosphate and nonphosphate buffers on thermolability of unstable hemoglobins.
Nature.
225
1970
1041
1042
30
Talbot
D
Collis
P
Antoniou
M
Vidal
M
Grosveld
F
Greaves
DR
A dominant control region from human β-globin locus conferring integration site-independent gene expression.
Nature.
33
1989
352
355
31
Pastzy
C
Mohandis
N
Stevens
ME
et al
Lethal α-thalassemia created by gene targeting in mice and its genetic rescue.
Nat Gene.
11
1995
33
39
32
Ciavatta
DJ
Ryan
TM
Farmer
SC
Townes
TM
Mouse model of human beta(o) thalassemia: targeted deletion of the mouse beta maj- and beta min-globin genes in embryonic stem cells.
Proc Natl Acad Sci U S A.
92
1995
9259
9263
33
Albitar
M
Cash
F
Peschle
C
Liebhaber
S
Developmental switch in the relative expression of the α1- and α2-globin genes in humans and in transgenic mice.
Blood.
79
1992
2471
2474
34
Kazazian
H
Woodhead
A
Hemoglobin A synthesis in the developing fetus.
N Engl J Med.
289
1973
58
62
35
Wood
W
Weatherall
D
Hemoglobin synthesis during human foetal development.
Nature.
244
1973
162
167
36
Peschle
C
Mavilio
F
Care
F
et al
Haemoglobin switching in human embryos: asynchrony of zeta-α and epsilon-γ globin switches in primitive and definitive erythropoietic lineage.
Nature.
313
1985
235
238
37
Albitar
M
Peschle
C
Liebhaber
S
Theta, zeta, and epsilon globin messenger RNAs are expressed in adults.
Blood.
74
1989
629
637
38
Luo
H
Xl
L
Frye
C
et al
Embryonic hemoglobins are expressed in definitive cells.
Blood.
94
1999
359
361
39
Chui
DHK
Wong
SC
Chung
S-W
Patterson
M
Bhargava
S
Poon
M-C
Embryonic zeta-globin chains in adults: a marker for alpha-thalassemia-1 haplotype due to a >17.5 kb deletion.
N Engl J Med.
314
1986
76
79
40
Blum
N
DeGrouchy
J
Alison
F
Deux cas de trisomie 13 avec persistance d'hemoglobine embryonnaire Gower 2 [Two cases of trimomy 13 with persistence of the embryonal hemoglobin Gower 2].
Annales de Genetique.
10
1967
138
140
41
Huehns
E
Hecht
F
Keil
J
Motulsky
A
Developmental hemoglobin anomalies in a chromosomal triplication: D1 trisomy syndrome.
Proc Natl Acad Sci U S A.
51
1964
89
97
42
Jensen
M
Murken
J
Hemoglobin chain synthesis in two children with trisomy 13: evidence for temporary imbalance during switch from gamma to beta chain synthesis.
Eur J Pediatr.
122
1976
151
157
43
Bethlenfalvay
N
Lourp
J
Greer
H
Translocation trisomy D syndrome 46, XX, D-, t(DqDq)+: report of a case with a note on the cold-instability of Hb Gower-2.
Pediatrics.
50
1972
928
930
44
Hill
AV
The possible effects of the aggregation of the molecules of haemoglobin on its dissociation curve.
J Physiol.
40
1910
iv
45
Fermi
G
Perutz
M
Haemoglobin and myoglobin.
Atlas of molecular structures in biology.
Phillips
DC
Richards
FM
1981
Clarendon Press
Oxford
46
Perutz
MF
Kilmartin
JV
Nishikura
K
Fogg
JH
Butler
PJG
Rollema
HS
Identification of residues contributing to the Bohr effect of human haemoglobin.
J Mol Biol.
138
1980
649
668
47
Kilmartin
JV
Fogg
JH
Perutz
MF
Role of C-terminal histidine in the alkaline Bohr effect of human hemoglobin.
Biochemistry.
19
1980
3189
3193
48
Matsukawa
S
Itatani
Y
Mawatari
K
Shimokawa
Y
Yoneyama
Y
Quantitative evaluation for the role of beta 146 His and beta 143 His residues in the Bohr effect of human hemoglobin in the presence of 0.1 M chloride ion.
J Biol Chem.
259
1984
11479
11486
49
Arnone
A
X-ray diffraction study of binding of 2,3-diphosphoglycerate to human deoxyhaemoglobin.
Nature.
237
1972
146
149
50
Gupta
RK
Benovic
JL
Rose
ZB
Location of the allosteric site for 2,3-bisphosphoglycerate on human oxy- and deoxyhemoglobin as observed by magnetic resonance spectroscopy.
J Biol Chem.
254
1979
8250
8255
51
Nishikura
K
Identification of histidine-122 α in human haemoglobin as one of the unknown alkaline Bohr groups by hydrogen-tritium exchange.
Biochem J.
173
1978
651
657
52
Charache
S
Jacobson
R
Brimhall
B
et al
Hb Potomac (101 Glu-Asp): speculations on placental oxygen transport in carriers of high-affinity hemoglobins.
Blood.
51
1978
331
338
53
Charache
S
Catalano
P
Burns
S
et al
Pregnancy in carriers of high-affinity hemoglobins.
Blood.
65
1985
713
718
54
Jia
L
Bonaventura
C
Bonaventura
J
Stamler
JS
S-nitrosohaemoglobin: a dynamic activity of blood involved in vascular control.
Nature.
380
1996
221
226
55
Lamb
P
Watt
P
Proudfoot
NJ
Negative regulation of the human embryonic globin genes zeta and epsilon.
Hemoglobin switching.
Stamatoyannopoulous
G
Nienhuis
AW
1988
269
277
Alan R. Liss
New York
56
Raich
N
Enver
T
Nakamoto
B
Josephson
B
Papayannopoulou
T
Stamatoyannopoulous
G
Autonomous developmental control of human embryonic globin gene switching in transgenic mice.
Science.
23
1990
1147
1149
57
Trepicchio
WL
Dyer
MA
Baron
MH
Developmental regulation of the human embryonic beta-like globin gene is mediated by synergistic interactions among multiple tissue- and stage-specific elements.
Mol Cell Biol.
13
1993
7457
7468
58
Vyas
P
Sharpe
JA
Watt
P
Higgs
DR
Wood
WG
Regulation of human embryonic globin genes zeta 2 and epsilon in stably transformed mouse erythroleukemia cells.
Blood.
80
1992
1832
1837
59
Hajjar
FM
Pearson
HA
Pharmacologic treatment of thalassemia intermedia with hydroxyurea.
J Pediatr.
125
1994
490
492
60
Sher
GD
Ginder
GD
Little
J
Yang
S
Dover
GJ
Olivieri
NF
Extended therapy with intravenous arginine butyrate in patients with beta hemoglobinopathies.
N Engl J Med.
332
1995
1606
1610

Author notes

J. Eric Russell, Abramson Research Bldg, Rm 316F, The Children's Hospital of Philadelphia, 34th St and Civic Center Blvd, Philadelphia, PA 19104; e-mail: jeruss@mail.med.upenn.edu.

Sign in via your Institution